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 24 毫秒
1.
Light-scattering and related studies on the polymerization behavior of the protein from the PM2 strain of TMV show that in phosphate buffer of ionic strength 0.1, the maximum extent of temperature-mediated polymerization occurs at pH values lower than in the case of TMV protein. The pH range of temperature-induced polymerization is from 5.0 to 6.0, contrasted with 5.0 to 7.5 for TMV protein. Velocity sedimentation studies show that PM2 protein at room temperature in phosphate buffer (I = 0.1) has sedimentation coefficients of 174 S, 104 S, and 4.3 S at pH values of 4.89, 5.53, and 7.5. Electron microscope studies show that at room temperature in phosphate buffer of 0.1 ionic strength at pH 5.53, PM2 protein has structures resembling essentially that of stacked double discs with an occasional helical structure. Similar studies of PM2 protein in 0.1 M ammonium acetate buffer at pH 5.2 show single, double, and double-double helices.  相似文献   

2.
3.
The heat effects accompanying the isothermalin vitro polymerization ofPr. mirabilis flagellin on short flagella fragments (seeds) have been measured in phosphate buffer pH 7, at various temperatures employing a batch microcalorimeter. Additionally, at 20 ?C, measurements have been performed in phosphate as well as Tris- HCl buffer at pH 7.5. The rate of both heat uptake and release during the process of polymerization was shown to be proportional to the rate of molar ellipticity changes observed by parallel circular dichroism experiments. No change in the state of protonation of flagellin occurs during the polymerization as indicated by the constancy of the enthalpy values determined in buffers with different heats of ionization. The apparent molar enthalpy of polymerization at 25 ?C, pH 7, is ?34.7±3 kcal per mole of flagellin, the relatively large error mainly resulting from uncertainties of the determination of the percentage of unpolymerized monomers after completion of the reaction. The most prominent feature of the results obtained in this study is the large temperature variation of the enthalpy, corresponding to a temperature independent heat capacity change ofδc p =?3039±100 cal per degree per mole of flagellin, the error limits referring to the standard deviation in a linear regression analysis.  相似文献   

4.
Osmotic pressure studies were carried on tobacco mosaic virus (TMV) and its components, protein and RNA, as well as on bis(3,3′-aminopropyl)amine, reported to be present in TMV preparations. Solvents were phosphate and barbital buffers at different values of pH and ionic strength. Measurements were made at room temperature. The Donnan effect was exhibited by TMV protein in phosphate buffer of 0.01 ionic strength at pH values ranging between 5.8 and 7.5. The observed values of the Donnan effect at pH 5.8 and 5.97 were in reasonable agreement with theoretical values calculated from the charge obtained by hydrogen ion titration. TMV-RNA in phosphate buffer at pH 7.5 and ionic strength 0.01 did not exhibit more than 1% of the expected Donnan effect. This is explained tentatively as the result of firm binding of metal ions. Negative values of osmotic pressure were observed with bis(3,3′-aminopropyl)amine. Similar anomalous osmosis was sometimes observed with TMV protein and with TMV. In agreement with earlier observations, TMV did not exhibit the Donnan effect in phosphate buffer of 0.01 ionic strength at pH values ranging from 5.5 to 8.0. However, TMV dialysed extensively in the presence of EDTA at pH 8.5 and TMV produced by reconstitution of purified protein and RNA did exhibit the Donnan effect in both phosphate and barbital buffers. The magnitude was of the same order as that calculated from the net charge determined by hydrogen ion titration. When reconstituted TMV, which did exhibit Donnan effect, was treated with calcium ions, the effect was abolished.  相似文献   

5.
d-Gluconate uptake was studied in whole cells of Arthrobacter pyridinolis; the uptake activity was inducible, mutable and showed saturation kinetics (Km = 5 μM). Uptake of d-gluconate was not mediated by a phosphoenolpyruvate: hexose phosphotransferase system, nor was it directly energized by ATP. A transmembrane pH gradient, ΔpH, of ?63 mV was generated by A. pyridinolis cells at pH 6.5, while at pH 7.5, ΔpH = 0. Addition of 8 μM d-gluconate significantly reduced the ΔpH. The transmembrane electrical potential, Δψ, which was ?87 mV over a range of pH from 5.5 to 7.5, was unaffected by the presence of substrate. d-Gluconate accumulated at the same rate and as the protonated solute, at both pH 6.5 and 7.5. Experiments in which a diffusion potential was generated in cyanide-treated cells, indicated that the Δψ did not energize transport. Rather, the rate of d-gluconate uptake correlated with and appeared to be determined by the rate of d-gluconate metabolism: (a) treatment of cells with valinomycin or nigericin, under conditions in which there was a loss of intracellular potassium, inhibited both d-gluconate uptake and the metabolism of pre-accumulated d-gluconate; (b) the effects of cyanide and azide on d-gluconate uptake were much more severe at pH 6.5 than pH 7.5, a pattern which paralleled the effects of these inhibitors on d-gluconate metabolism; (c) extraction and chromatography of intracellular label from d-gluconate uptake revealed that accumulation of unaltered d-gluconate was negligible; (d) a series of mutant strains with lower d-gluconate kinase activities also exhibited low rates of d-gluconate uptake; (e) spontaneous revertants of these mutant strains consistently regained both d-gluconate kinase activity and wild type levels of uptake.  相似文献   

6.
The entropy-driven polymerization of tobacco mosaic virus protein is favored by an increase in ionic strength, μ, and by a decrease in pH. The effect of ionic strength is interpreted in terms of salting-out and electrical work, a function of charge and, therefore, of pH as well as of μ. The extent of polymerization is measured in terms of a characteristic temperature, T1, corresponding to a characteristic value of the equilibrium constant, KcT1 is measured at an early stage in the polymerization process where the optical density increment from light scatter is 0.01. The theory developed encompassing both salting-out and electrical work terms relates 1T1 to μ approximately according to the equation, 1T1 = C + Bμ ? Aμ12, where C is the ratio of entropy to enthalpy, B is proportional to the salting-out constant divided by enthalpy, and Aμ12 depends upon the square of the charge and is proportional to the electrical work contribution divided by the enthalpy. Data in which μ varied from 0.025 to 0.150 at three pH values, 5.95, 6.35, and 6.50, were fitted to this equation and the parameters C, B, and A were evaluated. Experiments were also carried out at a constant μ of 0.10 at pH values in increments of 0.1 between 5.9 and 6.8. The theory predicts that, at constant μ, 1T1, corrected for the electrical work contribution, is a linear function of pH with a negative slope proportional to the number of hydrogen ions bound per protein unit during polymerization, divided by the enthalpy. The data obtained fit two straight lines with different slopes above and below pH 6.3. Independent experiments carried out by the method of Stevens and Loga show that the number of hydrogen ions bound per protein unit also differs above and below pH 6.3 and the ratio of these is the same as the ratio of the above mentioned slopes. The data, therefore, make it possible to evaluate the enthalpy to be 24.8 kcal/mol of associating A protein and, with this value, the parameters C, B, and A can be interpreted. Standard entropies range from 86 e.u. at pH 6.5 to 88.5 at pH 5.95 and the salting-out constant, KS, is 2.2 at all pH values studied. At μ = 0.10, the values of the electrical work contribution at pH 5.95, 6.35, and 6.50 are +0.298, +0.455, and +0.534 kcal/mol, respectively. Theoretical calculations from models predict values in agreement within a factor of less than two.  相似文献   

7.
Crystals of calotropin DI (Mr 23,400), have been prepared by microdialysis against 5% (w/v) polyethylene glycol 20,000 in water, pH 7.0. They have orthorhombic space group P212121 with cell parameters a = 57.5 A?, b = 86.2 A?, c = 40.3 A?. Crystals of calotropin DII (Mr 24,000), prepared by the same technique against 5% (w/v) polyethylene glycol 20,000 in phosphate buffer of low ionic strength, pH 7.0, display monoclinic space group C2 with cell parameters a = 135.8 A?, b = 32.0 A?, c = 47.7 A?, β = 103.80 °. In both cases, there is only one molecule in the asymmetric unit.  相似文献   

8.
The effects of absolute temperature (T), ionic strength (μ), and pH on the polymerization of tobacco mosaic virus protein from the 4 S form (A) to the 20 S form (D) were investigated by the method of sedimentation velocity. The loading concentration in grams per liter (C) was determined at which a just-detectable concentration (β) of 20 S material appeared. It was demonstrated experimentally that under the conditions employed herein, an equilibrium concentration of 20 S material was achieved in 3 h at the temperature of the experiment and that 20 S material dissociated again in 4 h or less to 4 S material either upon lowering the temperature or upon dilution. Thus, the use of thermodynamic equations for equilibrium processes was shown to be valid. The equation used to interpret the results, log (C?β) = constant + (ΔH12.3RT) + (ΔW1el2.3RT) ? K′ + ζpH, was derived from three separate models of the process, the only difference being in the anatomy of the constant; thus, the method of analysis is essentially independent of the model. ΔH1 and ΔW1el are the enthalpy and the change in electrical work per mole of A protein (the trimer of the polypeptide chain), Ks is the salting-out constant on the ionic strength basis, ζ is the number of moles of hydrogen ion bound per mole of A protein in the polymerization, and R is the gas constant. The three models leading to this equation are: a simple 11th-order equilibrium between A1 (the trimer of the polypeptide chain) and D, either the double disk or the double spiral of approximately the same molecular weight, designated model A; a second model, designated B, in which A1 was assumed to be in equilibrium with D at the same time that it is in equilibrium with A2, A3, etc., dimers and trimers, etc., of A1 in an isodesmic system; and a phase-separation model, designated model C, in which A protein is treated as a soluble material in equilibrium with D, considered as an insoluble phase. From electrical work theory, ΔWel1/T was shown to be essentially independent of T; therefore, in experiments at constant μ and constant pH the equation of log (C ? β) versus 1/T is linear with a slope of ΔH1/2.3R. The results fit such an equation over nearly a 20 °C-temperature range with a single value of ΔH1 of +32 kcal/mol A1. Results obtained when T and pH were held constant but μ was varied did not fit a straight line, which shows that more than simple salting-out is involved. When the effect of ionic strength on the electrical work contribution was considered in addition to salting-out, the data were interpreted to indicate a value of ΔW1el of 1.22 kcal/mol A1 at pH 6.7 and a value of 4.93 for Ks. When μ and T were held constant but pH was varied, and when allowance was made for the effect of pH changes on the electrical work contribution, a value of 1.1 was found for ζ. This means that something like 1.1 mol of hydrogen ion must be bound per mole of A1 protein in the formation of D. When this is added to the small amount of hydrogen ion bound per A1 before polymerization, at the pH values used, it turned out that for D to be formed, 1.5 H+ ions must be bound per A1 or 0.5 per protein polypeptide chain. This amounts to 1 H+ ion per polypeptide chain for half of the protein units, presumably those in one but not the other layer of the double disk or turn of the double spiral. When polymerization goes beyond the D stage, as shown by previously published data, additional H+ ions are bound. Simultaneous osmotic pressure studies and sedimentation studies were carried out, in both cases as a function of loading concentration C. These results were in complete disagreement with models A and C but agreed reasonably well with model B. The sedimentation studies permitted evaluation of the constant, β, to be 0.33 g/liter.  相似文献   

9.
The effect of the dipolar ions, glycine, glycylglycine, and glycylglycylglycine on the polymerization of tobacco mosaic virus (TMV) protein has been studied by the methods of light scattering and ultracentrifugation. All three dipolar ions promote polymerization. The major reaction in the early stage is transition from the 4 S to the 20 S state. As in the absence of dipolar ions, the polymerization is enhanced by an increase in temperature; it is endothermic and therefore entropy-driven. The effect of the dipolar ions can be understood in terms of their action as salting-out agents; they increase the activity coefficient of TMV A protein, the 4 S material, and thus shift the equilibrium toward the 20 S state. The salting-out constants, K, for the reaction in 0.10 ionic strength phosphate buffer at pH 6.7 was found by the light scattering method to be 1.6 for glycine, 2.5 for glycylglycine, and 2.5 for glycylglycylglycine. A value of 2.7 was obtained by the ultracentrifugation method for glycylglycine in phosphate buffer at 0.1 ionic strength and pH 6.8 at 10 degrees C. For both glycine and glycylglycine, K increases when the ionic strength of the phosphate buffer is decreased. This result suggests that electrolytes decrease the activity coefficient of the dipolar ions, a salting-in phenomenon. However, the salting-in constants evaluated from these results are substantially higher than those previously determined by solubility measurements. The effect of glycine and glycylglycine on polymerization was studied at pH values between 6.2 and 6.8. The effectiveness of both dipolar ions is approximately 50% greater at pH 6.8 than at pH 6.2. The variation of the extent of polymerization with pH in the presence of the dipolar ions is consistent with the interpretation that approximately one hydrogen ion is bound for half of the polypeptide units in the polymerized A protein.  相似文献   

10.
An in vivo assay of nitrate reductase activity was developed by vacuum infiltration of leaf discs or sections with a solution of 0.2 m KNO3 (with or without phosphate buffer, pH 7.5) and incubation of the infiltrated tissue and medium under essentially anaerobic conditions in the dark. Nitrite production, for computing enzyme activity, was determined on aliquots of the incubation media, removed at intervals.  相似文献   

11.
Homogenate preparations of human liver have been prepared and over 75% of the particulate neuraminidase activity (which comprises approx. 90% of the total activity) has been solubilized using 0.85% (w/v) Triton X-100 in 25 mM phosphate buffer (pH 6.8). The solubilized neuraminidase activity is extremely labile, but can be stabilized for at least 4 weeks at 2–4°C, using 10 mM N-acetylneuraminic acid. Kinetic characterization of homogenate and solubilized supernatant fluid neuraminidase activities indicated comparable pH optimum curves (maximum activity at pH 4.5–4.7) and apparent Km values (0.2–0.4 mM) for the synthetic fluorometric substrate 4-methylbelliferyl-α-D-N-acetylneuraminic acid. Isoelectric focusing has been performed on human liver homogenates and Triton X-100-solubilized neuraminidase activities, and the presence of several forms (4–6) with isoelectric points (pI values) between 4.4 and 5.2 has been demonstrated in both preparations. The similar kinetic and isoelectric focusing properties of the two preparations suggest that the solubilized enzyme activity is representative of the homogenate activity and that the solubilized enzyme is suitable for purification purposes.  相似文献   

12.
A new approach has been developed to monitor protein folding by utilizing osmotic pressure and a range of salt concentrations in a well characterized protein, bovine serum albumin (BSA). It is hypothesized that both the ‘effective’ osmotic molecular weight, Ae, and the solute/solvent interaction parameter, I, in the empirical relation MsolventMsolute = (RTϱAe)1gp + I [1] can be used as measures of protein folding. I is a measure of solvent perturbed by the solute and is thought to depend directly upon the solvent accessible surface area (ASA). It is reasoned that larger solvent accessible surface area of an unfolded or denatured protein should perturb more water and produce larger I-values. Thus I-values allow calculation of a unfolded protein fraction, fua, due to changes in relative solvent accessible surface area. It has been observed that Ae decreases for filamentous, denatured proteins due to segmental motion of the molecule [2]. This allows calculation of unfolded protein fraction from the effective molecular weight, fum. Colloid osmotic pressure of BSA was measured in a range of salt concentrations at 25°C, and pH = 7 (above the isoelectric point of BSA at pH = 5.4). Both S and I were used to monitor protein folding as the salt concentration was varied. In general, larger and variable I-values and smaller Ae were observed at salt concentrations less than 50 mmolal NaCl (Imax = 8.9), while constant I = 4.1 and Ae = 66,500 were observed above 50 mmolal NaCl. The two expressions for fractional unfolding (fua and fum) are in general agreement. Small differences in the parameters below 50 mmolal salt concentration are explained with well known shifts in the relative amounts of α-helix, β-sheet and random coil in denatured BSA. The relative amounts of these shifts agree with predictions in the literature attributed to continuous BSA expansion rather than an ‘all-or-none’ conversion.  相似文献   

13.
We have shown the dependence of the deamidation half-times of the peptides, GlyLeuGlnAlaGly and GlyArgGlnAlaGly upon pH, temperature, and ionic strength. Increase in temperature or ionic strength, variation of pH to pH′s higher or lower than pH 6, and the use of phosphate buffer rather than Tris buffer at high pH all decrease the half-time of dcamidation. Temperature increase of 20°C or pH change of 2 pH units decreases the half-time about fivefold, while increase of one ionic strength unit decreases the half-time about twofold. In pH 7.4, I = 0.2, 37.0°C phosphate buffer, the deamidation half-times are 663 ± 74 and 389 ± 56 days respectively for the two peptides, GlyLeuGlnAlaGly and GlyArgGlnAlaGly.These experiments should serve as a warning to peptide and protein experimenters that even the more stable glutaminyl residues are unstable with respect to deamidation in certain solvent conditions. These experiments also provide, along with previously reported experiments on asparaginyl peptides (7), some quantitative data to help with the extrapolation of in vitro deamidation experiments to in vivo deamidation conditions.  相似文献   

14.
The bivalve mollusc Scapharca inaequivalvis contains in the coelomic fluid erythrocytes with a dimeric (HbI) and a tetrameric (HbII) hemoglobin like the other members of the arcid family. The tetrameric protein is made up from two types of polypeptide chain, while the dimeric protein is made from a single type of chain which differs from the other two in terms of molecular weight and isoelectric point.The optical and circular dichroism spectra show that the heme environment in HbI and HbII resembles that of vertebrate hemoglobins, although distinctive features are present in the deoxygenated derivative. p]The dimeric HbI in the pH range 6 to 9 does not change its association state upon deoxygenation, while the tetrameric HbII polymerizes as indicated by the appearance of a fast peak in the sedimentation velocity patterns. The dependence of the areas and sedimentation coefficients of the fast and slow peaks on protein concentration is characteristic of a rapidly established association-dissociation equilibrium between tetramers and polymers higher than octamers. The pH, ionic strength and temperature dependence of polymer formation indicate that both hydrophobic and ionic interactions stabilize the polymers.The functional properties of HbI and HbII differ. HbI shows co-operative oxygen binding (h = 1·5) and a constant oxygen affinity (p12 = 7.8 mm Hg) over the pH range 5.5 to 9.5. HbII likewise shows co-operativity in oxygen binding (h = 2·0). Its oxygen affinity at neutral and alkaline pH values is slightly lower (p12 = 9.1 mm Hg) than that of the dimeric protein, but becomes higher at pH values below 6.5 due to the presence of an acid Bohr effect. At high protein concentrations, under conditions of extensive polymerization of the deoxygenated derivative, the oxygen affinity is lowered and co-operativity slightly increased. Both phenomena require that the oxygen affinity of the polymer be lower than that of the tetramer, consistent with the predictions of linkage theory.  相似文献   

15.
Crystals of fragment 1 of bovine prothrombin grown from phosphate at pH 7.5 are tetragonal, space group P41222 or P43212 with a = b = 79.5 A?, c = 84.9 A?, with probably one molecule of 22,000 molecular weight in the asymmetric unit. The presence of 17.5% carbohydrate in the fragment may account for the high liquid content (60%) of the crystals.  相似文献   

16.
The effect of phosphate buffer on the activity of jack bean urease was studied in the range of pH 5.80–8.07. The inhibition constants of phosphate buffer were determined by measuring initial reaction rates at each pH for a series of buffer concentrations at a series of urea concentrations. It was shown that: (1) at pH 5.80–7.49 the buffer is a competitive inhibitor of the enzyme with Ki,buffer increasing from 0.54 mM for pH 5.80 to 362 mM for pH 7.49, (2) the values of pKi,buffer are pH-dependent exhibiting a slope of −1 at pH 5.80–6.5 and a slope of −2 at pH 6.5–7.49, (3) from pH 7.62 as the pH is further raised the competitive inhibition of urease by the buffer was not observed, (4) the true competitive inhibitor of urease is H2PO4 ion, and (5) pH 6.5 and 7.6 correspond to the ionization constants of the active site groups of urease responsible for the inhibitory strength of H2PO4 ion.  相似文献   

17.
It was shown that a reversible endothermic association of TMV protein subunits (A protein) can take place at pH values below the isoelectric point as well as at pH 6.5. The polymerization occurring below the isoelectric point was found to be more complex than that at pH 6.5 probably because products other than the usual TMV-like rods were formed in addition to those rods and also because side-to-side aggregation of the rods took place readily. Kinetic studies indicated that polymerization can be treated as a second-order linear condensation. The rate of polymerization was found to be a critical function of pH, having a maximum value near pH 4.3. This behavior is at variance with the hypothesis that hydrogen-bonded carboxyl pairs play a dominant rate-determining role in the association of subunits. The dependence of the rate on pH was interpreted to indicate that electrostatic forces between subunits are a significant controlling factor in the polymerization of TMV protein.  相似文献   

18.
Escherichia coli has four hydrogenases (Hyd), three genes of which are encoded by the hya, hyb, and hyc operons. The proton-reducing and hydrogen-oxidizing activities of Hyd-2 (hyb) were analyzed in whole cells grown to stationary phase and cell extracts, respectively, during glycerol fermentation using novel double mutants. H2 production rate at pH 7.5 was decreased by ~3.5- and ~7-fold in hya and hyc (HDK 103) or hyb and hyc (HDK 203) operon double mutants, respectively, compared with the wild type. At pH 6.5, H2 production decreased by ~2- and ~5-fold in HDK103 and HDK203, respectively, compared with the wild type. At pH 5.5, H2 production was reduced by ~4.5-fold in the mutants compared with the wild type. The total hydrogen-oxidizing activity was shown to depend on the pH of the growth medium in agreement with previous findings and was significantly reduced in the HDK103 or HDK203 mutants. At pH 7.5, Hyd-2 activity was 0.26 U (mg protein)?1 and Hyd-1 activity was 0.1 U (mg protein)?1. As the pH of the growth medium decreased to 6.5, Hyd-2 activity was 0.16 U (mg protein)?1, and Hyd-1 was absent. Surprisingly, at pH 5.5, there was an increase in Hyd-2 activity (0.33 U mg protein)?1 but not in that of Hyd-1. These findings show a major contribution of Hyd-2 to H2 production during glycerol fermentation that resulted from altered metabolism which surprisingly influenced proton reduction.  相似文献   

19.
Isolation and characterization of isocitrate lyase of castor endosperm   总被引:1,自引:0,他引:1  
Isocitrate lyase (threo-DS-isocitrate glyoxylate-lyase, EC 4.1.3.1) has been purified to homogeneity from castor endosperm. The enzyme is a tetrameric protein (molecular weight about 140,000; gel filtration) made up of apparently identical monomers (subunit molecular weight about 35,000; gel electrophoresis in the presence of sodium dodecyl sulfate). Thermal inactivation of purified enzyme at 40 and 45 °C shows a fast and a slow phase, each accounting for half of the intitial activity, consistent with the equation: At = A02 · e?k1t + A02 · e?k2t, where A0 and At are activities at time zero at t, and k1 and k2 are first-order rate constants for the fast and slow phases, respectively. The enzyme shows optimum activity at pH 7.2–7.3. Effect of [S]on enzyme activity at different pH values (6.0–7.5) suggests that the proton behaves formally as an “uncompetitive inhibitor.” A basic group of the enzyme (site) is protonated in this pH range in the presence of substrate only, with a pKa equal to 6.9. Successive dialysis against EDTA and phosphate buffer, pH 7.0, at 0 °C gives an enzymatically inactive protein. This protein shows kinetics of thermal inactivation identical to the untreated (native) enzyme. Full activity is restored on adding Mg2+ (5.0 mm) to a solution of this protein. Addition of Ba2+ or Mn2+ brings about partial recovery. Other metal ions are not effective.  相似文献   

20.
Protein of the tobacco mosaic virus mutant E66 has lysine replacing asparagine of the type strain, vulgare, at position 140. Thus, E66 protein should have one more positive or one less net negative charge than vulgare at pH 6 to 7. To investigate the effect of charge, a comparative study of the polymerization of E66 and vulgare proteins at pH 6.0, 6.2, 6.4, 6.6, and 6.8 at ionic strengths 0.15, 0.10, and 0.05 was made by turbidimetry. Polymerization of E66 protein always proceeded at a lower temperature than vulgare. However, the extent of polymerization was much lower in E66, especially at the higher ionic strengths. Sedimentation velocity results paralleled those from turbidity measurements in that E66 protein polymerizes at lower temperatures than vulgare; the 20 S component is more abundant in E66 protein. Osmotic pressure measurements also show that E66 protein is more polymerized than vulgare, especially at lower pH values. Hydrogen ion titrations of E66 protein were carried out from pH 8 to 5 and back to pH 8 in 0.10 m KCl at three temperatures, 4, 10, and 15 °C. These titrations were reversible when carried out slowly. The isoionic point is near pH 5; thus the charge at pH 7.5 is ?3. The reversible titration results were correlated with the aggregates present at the various pH values and temperatures, determined from the areas under the schlieren peaks in sedimentation velocity experiments. It is found that hydrogen ion binding at the three pH values is correlated with the disappearance of the smallest aggregates and is independent of the type of higher polymer formed. To investigate the effect of ionic strength and pH on the characteristic temperature corresponding to an optical density increment of 0.01 by the method used previously for vulgare, two sets of turbidity measurements were carried out. In the first one the ionic strength was changed from 0.025 to 0.15 in increments of 0.025 at pH 6.0 and 6.4. In the other set, the ionic strength was kept constant at 0.10 and the pH changed from 5.9 to 6.7 in increments of 0.1 pH units. When the analysis of these data was carried out, ΔH1 = 30 kcal/mol was obtained. For the salting out constant a value of 1.7 was found, compared to 2.2 for vulgare, a result consistent with the fact that E66 should be less hydrophobic than vulgare. The electrical work term ΔWel also turns out to be about one-half that for vulgare, which is expected from the lower net negative charge on E66 protein.  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