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.
The occurrence of non-collagenous matrix proteins was studied in samples of tracheal cartilage from steers of different ages. The amounts of the 148 kDa cartilage protein in 4M-guanidinium chloride extracts and in subsequent trypsin digests of the extraction residues were determined by radioimmunoassay. Surprisingly, the 148 kDa-protein antigenicity was not changed by tryptic digestion, even though the protein was extensively degraded. The amount of the 148 kDa protein increased dramatically with age, both in the guanidinium chloride extract and in the subsequent tryptic digest, and reached maximal values at about 3 and 8 years respectively. The increase of the guanidinium chloride-soluble pool preceded that of the trypsin-digestible pool, possibly indicating a metabolic relationship. The ratio between the trypsin-digestible and guanidinium chloride-soluble pools increased continuously, and at 12 years of age close to 90% of the total 148 kDa protein detected was insoluble in guanidinium chloride. At all ages, the bulk of the cartilage collagen was insoluble both to extraction with guanidinium chloride and to tryptic digestion. The decreasing extractability of the 148 kDa protein was therefore not secondary to changes in the solubility of the collagen network. Other cartilage proteins, such as the link proteins and the 36 kDa protein, showed much smaller quantitative variations of a different character.  相似文献   

2.
The molecular size of pig liver carboxylesterase has been investigated under a variety of conditions of pH and ionic strength. From equilibrium and velocity sedimentation at pH 4.0 and pH 7.5, and from chromatography on Sephadex G-200,we conclude that the monomeric molecular weight is similar to 65,000 daltons and that the enzyme associates to form trimers. Association equilibrium constants for the monomer-trimer system were estimated to be 0.02 1-2 g-2 at pH 4 (concentration-dependent molecular weight data) and 2 times 10-5 1-2g-2 at pH 7.5 (frontal gel chromatographic results). These studies were aided by comparisons of the properties of the pig liver enzyme with those of chicken liver carboxylesterase, which is shown to exhibit the velocity and equilibrium sedimentation characteristics of a homogeneous protein with molecular weight similar to 65,000. Studies of pig and chicken liver carboxylesterases in 6 M guanidinium chloride, 0.1 M in beta-mercaptoethanol, support the proposition that the monomeric species of these enzymes have molecular weights of similar to 65,000. On polyacrylamide gel electrophoresis in SDS, there is no evidence for a major species of molecular weight less than similar to 65,000 for the pig enzyme, but ca. 50 percent of the chicken esterase is dissociated into two species of molecular weight similar to 30,000.  相似文献   

3.
More than 30 years ago, Nozaki and Tanford reported that the pK values for several amino acids and simple substances in 6 M guanidinium chloride differed little from the corresponding values in low salt (Nozaki, Y., and C. Tanford. 1967. J. Am. Chem. Soc. 89:736-742). This puzzling and counter-intuitive result hinders attempts to understand and predict the proton uptake/release behavior of proteins in guanidinium chloride solutions, behavior which may determine whether the DeltaG(N-D) values obtained from guanidinium chloride-induced denaturation data can actually be interpreted as the Gibbs energy difference between the native and denatured states (Bolen, D. W., and M. Yang. 2000. Biochemistry. 39:15208-15216). We show in this work that the Nozaki-Tanford result can be traced back to the fact that glass-electrode pH meter readings in water/guanidinium chloride do not equal true pH values. We determine the correction factors required to convert pH meter readings in water/guanidinium chloride into true pH values and show that, when these corrections are applied, the effect of guanidinium chloride on the pK values of simple substances is found to be significant and similar to that of NaCl. The results reported here allow us to propose plausible guanidinium chloride concentration dependencies for the pK values of carboxylic acids in proteins and, on their basis, to reproduce qualitatively the proton uptake/release behavior for the native and denatured states of several proteins (ribonuclease A, alpha-chymotrypsin, staphylococcal nuclease) in guanidinium chloride solutions. Finally, the implications of the pH correction for the experimental characterization of protein folding energetics are briefly discussed.  相似文献   

4.
The effects of pH and ionic strength on the midpoint reduction potential (Emp) of Clostridium acidi-urici ferredoxin were determined using hydrogen gas and hydrogenase. The Emp of native ferredoxin at 24-25 degrees in 0.1 M Tris-chloride buffer, pH 7.0, is--0.434 V. In the pH range examined, the Emp becomes approximately 13 mv more negative per each pH unit increase. A plot of the log of ionic strength versus the apparent Emp of ferredoxin in 0.1 M Tris-chloride buffer, pH 7.5, Was linear over the range of 1.0 to 0.01 ionic strength with Emp values of--0.414 and--0.475 V, respectively, at these extremes. This effect is the same with sodium chloride, sodium bromide, or ammonium sulfate. Potassium phosphate buffer caused a similar change, but the absolute values of Emp differed from those obtained in the presence of the other salts. This effect of pH and ionic strength on Emp may be general for clostridial-type (Fe4S4)2-ferredoxins, since the apparent Emp of Clostridium pasteurianum ferredoxin is affected in a similar manner by these two variables. The Emp of this ferredoxin in 0.1 M Tris-chloride buffer pH 7.0, is--0.405 V. Since the NH2-terminal amino acid residue, Ala1, and Tyr2 of C. acidi urici ferredoxin are near an (Fe4S4)2-cluster in the protein, the apparent Emp of derivatives that contained amino acid replacements in these two positions were determined. Under similar conditions, the Emp of most of the 13 derivatives examined, including those of [Leu2]- and[3-NH2-Tyr30]ferredoxin, is approximately the same as that of native ferredoxin. However, the Emp of [His2]ferredoxin is approximately 15 mv more positive, whereas that of [Trp2]ferredoxin is 22 mv more negative than that of native C. acidi-urici ferredoxin. Variations in sodium chloride concentration and pH also affected the apparent Emp of the derivatives. It is suggested that the changes observed in the Emp of C. acidi-urici ferredoxin are caused by protein conformational changes.  相似文献   

5.
The molecular weights of alkylated small subunits of ribulose-1,5-bisphosphate carboxylase of pea and spinach were determined from gel filtration data in the presence of 6M guanidinium chloride as 12,800 and 13,500, respectively. In the presence of 0.1 M sodium phosphate (pH 12.0) these molecules chromatograph at the same position as chymotrypsinogen (M=25,700) on Sephadex G-75. The intrinsic viscosity of the small subunit of spinach ribulose bisphosphate carboxylase, measured in this solvent, was [η] = 30 cm3/gm, while the intrinsic viscosity of chymotrypsinogen measured in this solvent was [η] = 2.4 cm3/gm. These data rule out a globular dimeric model for the sructure of the small subunit in 0.1 M sodium phosphate (pH 12.0).  相似文献   

6.
Diffusion and partitioning of proteins in charged agarose gels.   总被引:4,自引:2,他引:2       下载免费PDF全文
The effects of electrostatic interactions on the diffusion and equilibrium partitioning of fluorescein-labeled proteins in charged gels were examined using fluorescence recovery after photobleaching and gel chromatography, respectively. Measurements were made with BSA, ovalbumin, and lactalbumin in SP-Sepharose (6% sulfated agarose), in phosphate buffers at pH 7 and ionic strengths ranging from 0.01 to 1.0 M. Diffusivities in individual gel beads (D) and in the adjacent bulk solution (D infinity) were determined from the spatial Fourier transform of the digitized two-dimensional fluorescence recovery images. Equilibrium partition coefficients (phi) were measured by recirculating protein solutions through a gel chromatography column until equilibrium was reached, and using a mass balance. Diffusion in the gel beads was hindered noticeably, with D/D infinity = 0.4-0.5 in each case. There were no effects of ionic strength on BSA diffusivities, but with the smaller proteins (ovalbumin and lactalbumin) D infinity increased slightly and D decreased at the lowest ionic strength. In contrast to the modest changes in diffusivity, there were marked effects of ionic strength on the partition coefficients of these proteins. We conclude that for diffusion of globular proteins through gel membranes of like charge, electrostatic effects on the effective diffusivity (Deff = phi D) are likely to result primarily from variations in phi with only small contributions from the intramembrane diffusivity.  相似文献   

7.
Proteoglycans were extracted, in a yield of about 90%, from costal cartilage of young, growing guinea-pigs. Three solvents were used in sequence: 0.4 M guanidine - HCl, pH 5.8, 4 M guanidine - HCl, pH 5.8, and 4 M guanidine - HCl/0.1 M EDTA, pH 5.8. The proteoglycans were purified and fractionated by cesium chloride density gradient ultracentrifugation under associative and dissociative conditions. Gel chromatography on Sepharose 2 B of proteoglycan fractions from associative centrifugations showed the presence of both aggregated and monomer proteoglycans. The ratio of aggregates to monomers was higher in the second extract than in the other two extracts. Dissociative gradient centrifugation gave a similar distribution for proteoglycans from all three extracts. Thus, with decreasing buoyant density there were decreasing ratios of polysaccharide to protein, and of chondroitin sulfate to keratan sulfate. In addition, there was with decreasing density an increasing ratio of chondroitin 4-sulfate to chondroitin 6-sulfate. Amino acid analyses of dissociative fractions were inaccordance with previously published results. On comparing proteoglycan monomers of the three extracts, significant differences were found. Proteoglycans, extracted at low ionic strength, contained lower proportions of protein, keratan sulfate, chondroitin 6-sulfate and basic amino acids than those of the second extract. The proteoglycans of the third extract also differed from those of the other extracts. The results indicate that the proteoglycans of guinea-pig costal cartilage exist as a very polydisperse and heterogenous population of molecules, exhibiting variations in aggregation capacity, molecular size, composition of protein core, degree of substitution of the protein core, as well as variability in the type of polysaccharides substituted.  相似文献   

8.
The structure of protein SI of Thermus thermophilus (M = 61 kDa) in solution at low and moderate ionic strengths (0 M and 100 mM NaCl, respectively) has been studied by small-angle X-ray and neutron scattering. It was found that protein S1 has a globular conformation under both ionic conditions. The modelling of different packing of six homologous domains of S1 on the basis of the NMR-resolved structure of one domain showed that the best fit of calculated scattering patterns from such complexes to experimental ones is observed at a compact package of the domains. The calculated value of the radius of gyration of the models is 28-29 angtroms, which is characteristic for globular proteins with a molecular mass of about 60 kDa. It was found that protein S1 has a tendency to form associates, and the type of the associate depends on ionic strength. These associates have, in general, two or three monomers at a moderate ionic strength, while at a low ionic strength the number of monomers exceeds three and they are packed in a compact manner. Strongly elongated associates were observed in neutron experiments at a moderate ionic strength in heavy water. The association of protein molecules was also confirmed by the data of dynamic light scattering. From these data, the translational diffusion coefficient of protein S1 at a moderate ionic strength was calculated to be (D20,w = (2.7 +/- 0.1) x 10(-7)cm2/s). This value is essentially smaller than the expected value (D20,w = (5.8 - 6.0) x 10(-7)cm2/s) for the S1 monomer in the globular conformation, indicating the association of protein molecules under equilibrium conditions.  相似文献   

9.
Solubility of lysozyme chloride was determined in the absence of added salt and in the presence of 0.05-1.2 M NaCl, starting from isoionic lysozyme, which was then brought to pH values from 9 to 3 by addition of HCl. The main observation is the absence of a salting-in region whatever the pH studied. This is explained by a predominant electrostatic screening of the positively charged protein and/or by adsorption of chloride ions by the protein. The solubility increases with the protein net charge at low ionic strength, but the reverse is observed at high ionic strength. The solubility of lysozyme chloride seems to become independent of ionic strength at pH approximately 9.5, which is interpreted as a shift of the isoionic pH (10.8) to an isoelectric pH due to chloride binding. The crystallization at very low ionic strength, where lysozyme crystallizes at supersaturation values as low as 1.1, amplifies the effect of pH on protein solubility. Understanding the effect of the net charge and of ionic strength on protein-protein interactions is valuable not only for protein crystal growth but more generally also for the formation of protein-protein or protein-ligand complexes.  相似文献   

10.
Streptococcal serum opacity factor (SOF) is a protein that clouds the plasma of multiple mammalian species by disrupting high density lipoprotein (HDL) structure. Intravenous infusion of low dose SOF (4 µg) into mice reduces their plasma cholesterol concentrations?~?40% in 3 h. Here we investigated the effects of pH, ionic strength, temperature, and denaturation with guanidinium chloride (GdmCl) on SOF stability and its reaction vs HDL. SOF stability was tested by pre-incubation of SOF at various temperatures, pH’s, and GdmCl concentrations and measuring the SOF reaction rate at pH 7.4 and 37?°C. SOF retained activity at temperatures up to 58?°C, at pH 4 to 10, and in 8.5 M GdmCl after being returned to standard buffer conditions. The effects of GdmCl, pH, and ionic strength on the SOF reaction rates were also measured. SOF was inactive at GdmCl?≥?1 M; SOF was most active at pH 5, near its isoelectric point and at an ionic strength of 3 (in NaCl). These data reveal that SOF is a stable protein and suggest that its activity is determined, in part, by the effects of pH and ionic strength on its overall charge relative to that of its reaction target, HDL.  相似文献   

11.
The far-ultraviolet circular dichroic spectrum of the 39-residue peptide hormone porcine corticotropin and the biologically active fragment corticotropin 1–24 is negative from 250 nm to 195 nm in water, but in 6M guanidinium chloride a positive band appears at about 225 nm. The temperature and guanidinium chloride dependence of this spectral transition indicates the absence of any stable ordered secondary structure in corticotropin and the spectrum is seen to be in only partial agreement with results using the model peptide chromophore, Ala-Ala-Ala. Using oligopeptides containing aromatic amino acid residues sandwiched between glycyl residues, it is shown that the shape and intensity of the corticotropin 225 nm positive band which appears in 6M guanidinium chloride is in agreement with the far-ultraviolet transitions of the aromatic chromophores in the hormone. Curve resolution of the near-ultraviolet circular dichroic spectrum of corticotropin and comparison of the rotational strengths of the phenylalanyl and tyrosyl bands reveals no evidence for increased rotational freedom in 6M guanidinium hydrochloride. Spectral changes are observed, however, in the transitions arising from the single tryptophan. This study suggests that corticotropin in aqueous solution may serve as a better model for the circular dichroic spectrum of the aperiodic regions in globular proteins than either synthetic homopolypeptides or reference proteins for which spectral and X-ray diffraction data are available.  相似文献   

12.
The interactions of partially unfolded proteins provide insight into protein folding and protein aggregation. In this work, we studied partially unfolded hen egg lysozyme interactions in solutions containing up to 7 M guanidinium chloride (GdnHCl). The osmotic second virial coefficient (B(22)) of lysozyme was measured using static light scattering in GdnHCl aqueous solutions at 20 degrees C and pH 4.5. B(22) is positive in all solutions, indicating repulsive protein-protein interactions. At low GdnHCl concentrations, B(22) decreases with rising ionic strength: in the absence of GdnHCl, B(22) is 1.1 x 10(-3) mLmol/g(2), decreasing to 3.0 x 10(-5) mLmol/g(2) in the presence of 1 M GdnHCl. Lysozyme unfolds in solutions at GdnHCl concentrations higher than 3 M. Under such conditions, B(22) increases with ionic strength, reaching 8.0 x 10(-4) mLmol/g(2) at 6.5 M GdnHCl. Protein-protein hydrodynamic interactions were evaluated from concentration-dependent diffusivity measurements, obtained from dynamic light scattering. At moderate GdnHCl concentrations, lysozyme interparticle interactions are least repulsive and hydrodynamic interactions are least attractive. The lysozyme hydrodynamic radius was calculated from infinite-dilution diffusivity and did not change significantly during protein unfolding. Our results contribute toward better understanding of protein interactions of partially unfolded states in the presence of a denaturant; they may be helpful for the design of protein refolding processes that avoid protein aggregation.  相似文献   

13.
T Pawelczyk  S Angielski 《Biochimie》1992,74(2):171-176
The effects of changing ionic strength on the activity of the 2-oxoglutarate dehydrogenase complex from pig kidney cortex were explored. This enzyme complex is found to be influenced in many ways by the ionic strength of the reaction medium. The enzyme shows an optimum activity at 0.1 M ionic strength. Increase in ionic strength from 0.1 M to 0.2 M resulted in a decrease of S0.5 for 2-oxoglutarate, and in an increase of S0.5 for NAD. Changes in ionic strength over the range of 0.05-0.2 M have little, if any, effect on S0.5 for CoA. The Hill coefficient for 2-oxoglutarate and NAD at 0.2 M ionic strength was 1.0, whereas at 0.05 M ionic strength it was 0.85 and 1.2 for 2-oxoglutarate and NAD, respectively. At 0.05 M ionic strength the pH optimum of the enzyme ranges between 7.4-7.6, but at 0.15 M ionic strength the pH optimum shifts to 7.8. The magnitude of inhibition of enzyme activity by ATP is not influenced by changes in ionic strength in the absence of calcium. However, in the presence of Ca2+, increases in ionic strength lower the inhibitory effects of ATP. The Si0.5 for ATP in both presence and absence of Ca2+ was not affected by changes in ionic strength in the range of 0.1-0.2 M. In contrast, the Sa0.5 for ADP in the absence of Ca2+ decreases as ionic strength increases. In the presence of calcium and 0.2 M ionic strength ADP has no effect on 2-oxoglutarate dehydrogenase complex activity.  相似文献   

14.
1. Human bronchial lysozyme was isolated from nonpurulent secretions and studied by circular dichroism (CD) spectroscopy for its conformational properties. 2. The two negative bands at 208 and 222 nm indicated that the peptide chain adopted an alpha-helical structure in physiological conditions. 3. The molecule was stable at pH 1.0 but not at pH 12.0. 4. Increasing ionic strength by adding NaCl up to 1 M did not change the CD spectra. 5. Complete unfolding of the molecule by guanidinium chloride was obtained only at the concentration of 6 M. 6. Bronchial lysozyme was also denatured by sodium dodecyl sulphate. 7. The molecule was stable when mild reduction was performed at 37 degrees C for 30 min but was completely unfolded after heating at 100 degrees C for 3 min.  相似文献   

15.
The three-dimensional structure of carbonic anhydrase is a result of specific folding of the protein chain to form a compact, globular molecule. Fluorescence measurements on the nature of the rate-limiting steps in folding from the random coil to the native structure show that each step involves an actual folding reaction of the protein chain. Emission intensity and polarization of the intrinsic fluorescence due to tryptophan residues reach a maximum during the early period of the folding process. The changes occur in at least three kinetic phases (tau1 less than 3 S, tau2 = 1 min, tau3 = 10 min, 1 M guanidinium chloride, 2 M NaC1, pH 7, 20 degrees C). None of these phases are explained by configurational changes in the fully unfolded chain. The results are consistent with a kinetic scheme that involves stepwise acquisition of the specific folded structure of the native enzyme.  相似文献   

16.
The binding of deoxycholate and dodecyl sulfate to the mouse and bovine myelin basic proteins and two peptide fragments, obtained by cleavage of the bovine basic protein at its single tryptophan residue, was examined. Complete equilibrium binding isotherms for both detergents were obtained by examining their binding to each of the polypeptides immobilized on agarose. The bulk of the binding of dodecyl sulfate was found to be highly cooperative, and at saturation all four polypeptides bound far more detergent than globular, water-soluble proteins. The sum of the dodecyl sulfate bound by each of the two bovine basic protein cleavage fragments was almost twice that bound by the intact protein at saturation, suggesting that cleavage of the bovine basic protein exposes sites for additional binding of dodecyl sulfate. At pH values below pH 8.0, an additional cooperative transition was observed below the critical micelle concentration of sodium dodecyl sulfate in the binding isotherms of all four polypeptides. The midpoint of this transition corresponded to an apparent pK of approximately 5.5; however, the destruction of 90% of the histidine residues in the bovine basic protein had no effect on this transition. At pH 9.2 and moderate ionic strength (I = 0.1), the bulk of the binding of deoxycholate to the mouse and bovine basic proteins occurred at and above the critical micelle concentration of the detergent; and saturation values of deoxycholate binding to these two proteins were considerably higher than that reported for globular, water-soluble proteins. In marked contrast to the results with dodecyl sulfate, neither cleavage fragment was observed to bind deoxycholate. The results suggest that the higher ordered structure of the bovine basic protein may play an important role in the binding of anionic amphiphiles to the protein.  相似文献   

17.
The solution structure of the histone-like DNA-binding protein, HBsu, from Bacillus subtilis in 2 mM sodium cacodylate, pH 7.5, is sensitive to the ionic strength of the buffer. This was shown by circular dichroism measurements at different concentrations of sodium chloride and potassium fluoride. The stability of HBsu is also influenced; at HBsu concentrations of about 0.1 mg.ml-1, melting temperatures of 32 degrees C and 55 degrees C were found in the absence of potassium fluoride and in the presence of 0.5 M potassium fluoride, respectively, exhibiting drastic ionic-strength-dependent differences in the temperature-induced unfolding of HBsu. Furthermore, at low ionic strength, circular dichroism spectra vary markedly depending on the HBsu concentration in the approximate range 0.2-3 mg.ml-1. Such protein-concentration-dependent differences in the spectra were not observed in the presence of 0.5 M potassium fluoride. Very similar circular dichroism spectra of HBsu and the histone-like DNA-binding protein of Bacillus stearothermophilus (HBst) at high ionic strength, indicate comparable structures of both proteins under these conditions. Estimation of the secondary structure content from the circular dichroism spectra yields data which are in satisfactory agreement with the values obtained from the crystal structure of HBst. Transition temperatures of 45 degrees C and 61 degrees C were found in differential scanning calorimetric measurements performed with HBsu in potassium-fluoride-free buffer and in the presence of 0.5 M potassium fluoride, respectively. The thermodynamic data point to the melting of native HBsu dimers into two denatured monomers.  相似文献   

18.
Small angle neutron scattering (SANS) results revealed that contrary to most reports C9 is not a globular protein. Its radius of gyration (Rg) at pH 8 and an ionic strength of 0.5 is 32.2 +/- 1.4 A increasing to 35 A at physiologic ionic strength. In contrast, C8, which has a 2.2-fold larger mass, has a similar Rg value [34.6 +/- 1.6 A]. Calibration plots of Rg vs. M(r) indicate that native C8 is a spherical protein whereas native C9 is elongated. From previous reports it was known that native C8 and C9 associate in solutions of low ionic strength. SANS results confirmed this observation but also demonstrated that C8-C9 heterodimers are already formed at physiologic ionic strength. The dimeric complex is globular [Rg = 40 +/- 0.8 A] indicating that the proteins associate side-by-side rather than end-to-end. In contrast, in presence of the drug Suramin, a potent inhibitor of the assembly of the C5b-9 complex, C9 forms a complex with twice the molecular mass that is still elongated (Rg = 48.8 +/- 0.8 A), suggesting that in this case the protein dimerizes end-to-end via a bridging Suramin molecule.  相似文献   

19.
1. The process of denaturation of the chicken muscle dimeric enzyme triosephosphate isomerase on addition of guanidinium chloride has been studied at pH 7.6, the pH at which the recovery of activity is optimal (100%) on removal of denaturant. Determinations of the sedimentation coefficient, intrinsic viscosity, molecular weight (by sedimentation equilibrium studies) and the absorption coefficient at 280 nm in various concentrations of guanidinium chloride concurred in showing a single, sharp transition at about 0.7 M guanidinium chloride at a protein concentration 1-5 mg/ml from the native enzyme to the dissociated, unfolded chains of the monomer. Relative fluorescent intensity measurements revealed a single transition at about 0.4 M guanidinium chloride at enzyme concentrations of about 0.05 mg/ml. 2. The process of denaturation in different guanidinium chloride concentrations was first order with respect to enzyme and about sixth order with respect to denaturant. 3. The rate of attainment of equilibrium during the renaturation obeyed second-order/first-order reversible kinetics. It was concluded that the rate-determining step in renaturation at pH 7.6 must be the association of two subunits.  相似文献   

20.
The thermal stability of the methionine repressor protein from Escherichia coli (MetJ) has been examined over a wide range of pH (pH 3.5-10) and ionic strength conditions using differential scanning calorimetry. Under reducing conditions, the transitions are fully reversible, and thermograms are characteristic of the cooperative unfolding of a globular protein with a molecular weight corresponding to the MetJ dimer, indicating that no dissociation of this dimeric protein occurs before unfolding of the polypeptide chains under most conditions. In the absence of reducing agent, repeated scans in the calorimeter show only partial reversibility, though the thermodynamic parameters derived from the first scans are comparable to those obtained under fully reversible conditions. The protein is maximally stable (Tm 58.5 degrees C) at about pH 6, close to the estimated isoelectric point, and stability is enhanced by increasing ionic strength in the range I = 0.01-0.4 M. The average calorimetric transition enthalpy (delta Hm) for the dimer is 505 +/- 28 kJ mol-1 under physiological conditions (pH 7, I = 0.125, Tm = 53.2 degrees C) and shows a small temperature dependence which is consistent with an apparent denaturational heat capacity change (delta Cp) of about +8.9 kJ K-1 mol-1. The effects of both pH and ionic strength on the transition temperature and free energy of MetJ unfolding are inconsistent with any single amino acid contribution and are more likely the result of more general electrostatic interactions, possibly including significant contributions from electrostatic repulsion between the like-charged monomers which can be modeled by a Debye-Hückel screened potential.  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