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.
Mitra RC  Zhang Z  Alexov E 《Proteins》2011,79(3):925-936
Protein-protein association is a pH-dependent process and thus the binding affinity depends on the local pH. In vivo the association occurs in a particular cellular compartment, where the individual monomers are supposed to meet and form a complex. Since the monomers and the complex exist in the same micro environment, it is plausible that they coevolved toward its properties, in particular, toward the characteristic subcellular pH. Here we show that the pH at which the monomers are most stable (pH-optimum) or the pH at which stability is almost pH-independent (pH-flat) of monomers are correlated with the pH-optimum of maximal affinity (pH-optimum of binding) or pH interval at which affinity is almost pH-independent (pH-flat of binding) of the complexes made of the corresponding monomers. The analysis of interfacial properties of protein complexes demonstrates that pH-dependent properties can be roughly estimated using the interface charge alone. In addition, we introduce a parameter beta, proportional to the square root of the absolute product of the net charges of monomers, and show that protein complexes characterized with small or very large beta tend to have neutral pH-optimum. Further more, protein complexes made of monomers carrying the same polarity net charge at neutral pH have either very low or very high pH-optimum of binding. These findings are used to propose empirical rule for predicting pH-optimum of binding provided that the amino acid compositions of the corresponding monomers are available.  相似文献   

2.
Saturation analysis of equilibrium binding of iodinated thyrotropin (125I-TSH) to normal human thyroid preparations yielded linear Scatchard plots under non-physiological conditions of pH 6.0 or 20 mM Tris/acetate buffer, pH 7.4. The apparent equilibrium dissociation constant of this binding was approximately 10(-8) M. By contrast, nonlinear plots were obtained under standard conditions of pH 7.4 and 40 mM Tris/acetate buffer. Resolution of the components of these curves by computer analysis revealed the presence of at least two classes of binding sites, one of which is of a low capacity and high affinity (approximately 10(-10) M) consistent with receptor binding. The other component is of a high capacity and lower affinity. Binding to non-target tissues of muscle, parathyroid, mammary carcinoma, and placenta was only demonstrable at pH 6.0 or in 20 mM Tris/acetate buffer, pH 7.4, yielding linear Scatchard plots with similar binding affinity (approximately 10(-8)M) to normal thyroid but much reduced capacity. Preincubation of thyroid tissue at 50 degrees C resulted in an apparent selective loss of the high affinity component of binding measured under standard conditions. Kinetic experiments on the dissociation of bound 125I-TSH were undertaken to determine whether the non-linearity of Scatchard plots was due to two or more classes of binding sites or negative cooperativity. It was found that the experimental determinant that is presently ascribed to a negative cooperativity phenomenon regulating receptor affinity (i.e. an enhanced dilution-induced dissociation rate in the presence of excess native hormone), although apparently hormone-specific, was demonstrated under nonphysiological binding conditions and in non-target tissue. Significantly, the phenomenon was found under conditions of pH 6.0 or 20 mM Tris where a linear Scatchard plot was obtained. The evidence thus suggests that 125I-TSH binds to heterogeneous binding sites (of which the high affinity is probably the receptor for TSH) and that the enhanced dilution-induced dissociation of bound hormone by native hormone for this system, is only a characteristic of the low affinity binding site (maybe gangliosides).  相似文献   

3.
The binding of CTP and ATP to aspartate transcarbamylase at pH 7.8 and 8.5 at 25 degrees has been investigated by equilibrium dialysis and flow microcalorimetry. The binding isotherms for CTP at both pH 7.8 and 8.5 and ATP AT PH 8.5 can be fit by a model which assumes three tight, three moderately tight, and six weak binding sites. The binding isotherms for ATP at pH 7.8 are best fit by a model which assumes six tight and six weaker sites. Both finite differenceH binding and finite differenceS binding are negative for both nucleotides at both pH values, so that the binding is enthalpy driven. For both nucleotides, finite differenceH is the same for the first two classes of binding sites, implying that the difference in the dissociation constants of these two classes of sites is the result of entropic effects. Direct pH measurements and calorimetric measurements in two buffers with very different heats of ionization (Tris and Hepes) indicate that the binding of both nucleotides is accompanied by the binding of protons. In the pH range 6.7-8.4, the number of moles of protons bound per mole of nucleotide increases as the pH decreases.  相似文献   

4.
The kinetics of the binding reaction of MS2 phage to free F pili, which were highly purified from Escherichia coli, has been studied using a membrane filter assay. The rate of dissociation (kd) of the MS2-phage--F-pilus complex is very slow and follows first-order kinetics with a half-life of 4.2 h at 30 degrees C in the standard buffer. The dissociation rate is rather insensitive to temperature, but becomes more rapid at high ionic strength or at basic pH. In a 0.25 M ionic strength buffer, the half-life of the complex is about 1.0 min. The rate of association is very fast and follows second-order kinetics with the rate constant for association (ka) being 8 x 10(7) M-1 s-1 at 30 degrees C in the standard buffer. The rate of association is almost insensitive to ionic strength but slightly sensitive to pH or temperature. Monovalent cations can also promote the binding reaction as well as divalent cations but the complex formed with monovalent cation is unstable. A study of the kinetics of dissociation suggests that there are two types of interaction between MS2 phage and F pilus; one is a strong interaction formed with divalent cations and the other is a weak one formed with monovalent cations. The physical nature of the bonds involved in the former and the latter seems to be mainly electrostatic and non-electrostatic respectively. The mechanism of the binding reaction is discussed.  相似文献   

5.
In this study the pH dependence of the thermal stability of Sso7d from Sulfolobus solfataricus is analyzed. This small globular protein of 63 residues shows a very marked dependence of thermal stability on pH: the denaturation temperature passes from 65.2 degrees C at pH 2.5 to 97.9 degrees C at pH 4.5. Analysis of the data points out that the binding of at least two protons is coupled to the thermal unfolding. By linking the proton binding to the conformational unfolding equilibrium, a thermodynamic model, which is able to describe the dependence upon the solution pH of both the excess heat capacity function and the denaturation Gibbs energy change for Sso7d, is developed. The decreased stability in very acid conditions is due to the binding of two protons on identical and noninteracting sites of the unfolded state. Actually, such sites are two carboxyl groups possessing very low pKa values in the native structure, probably involved in salt-bridges on the protein surface.  相似文献   

6.
Interaction of soybean beta-amylase with glucose   总被引:1,自引:0,他引:1  
The interaction of soybean beta-amylase with glucose was investigated by inhibition kinetics studies and spectroscopic measurements. The inhibition type, inhibitor constant (Ki) and dissociation constant (Kd) of beta-amylase-glucose complex were dependent on pH. At pH 8.0, glucose behaved as a competitive inhibitor (Ki = 34 mM). Binding of glucose produced a characteristic difference spectrum and a change of circular dichroism (CD) at pH 8.1. By using difference absorbance at 292 nm and difference ellipticity at 290 nm, Kd values for beta-amylase-glucose complex were determined to be 45 and 46 mM, respectively. In contrast to pH 8.0, glucose behaved as a mixed-type inhibitor (Ki = 320 mM) at pH 5.4. The Kd values obtained from the difference spectrum were increased by lowering the pH from 8. The pH dependence of the Ki and Kd values suggested that one ionizable group of pK = 8.0, which is shifted to 6.9 by the binding of glucose, controls the binding affinity of glucose. The binding of glucose competed with the binding of cyclohexaamylose and maltose at pH 8.0. The modification of SH groups of the enzyme affected the binding of glucose but did not affect the binding of maltose or cyclohexaamylose at pH 8.0. It was concluded from these results that the binding site of glucose is different from that of maltose and cyclohexaamylose. Presumably, glucose may bind to the subsite 1 of soybean beta-amylase.  相似文献   

7.
D E Hill  G G Hammes 《Biochemistry》1975,14(2):203-213
Equilibrium binding studies of the interaction of rabbit muscle phosphofructokinase with fructose 6-phosphate and fructose 1,6-bisphosphate have been carried out at 5 degrees in the presence of 1-10 mM potassium phosphate (pH 7.0 and 8.0), 5 mM citrate (pH 7.0), or 0.22 mm adenylyl imidodiphosphate (pH 7.0 and 8.0). The binding isotherms for both fructose 6-phosphate and fructose 1,6-bisphosphate exhibit negative cooperativity at pH 7.0 and 8.0 in the presence of 1-10 mM potassium phosphate at protein concentrations where the enzyme exists as a mixture of dimers and tetramers (pH 7.0) or as tetramers (pH 8.0) and at pH 7.0 in the presence of 5 mM citrate where the enzyme exists primarily as dimers. The enzyme binds 1 mol of either fructose phosphate/mol of enzyme monomer (molecular weight 80,000). When enzyme aggregation states smaller than the tetramer are present, the saturation of the enzyme with either ligand is paralleled by polymerization of the enzyme to tetramer, by an increase in enzymatic activity and by a quenching of the protein fluorescence. At protein concentrations where aggregates higher than the tetramer predominate, the fructose 1,6-bisphosphate binding isotherms are hyperbolic. These results can be quantitatively analyzed in terms of a model in which the dimer is associated with extreme negative cooperativity in binding the ligands, the tetramer is associated with less negative cooperativity, and aggregates larger than the tetramer are associated with little or no cooperativity in the binding process. Phosphate is a competitive inhibitor of the fructose phosphate sites at both pH 7.0 and 8.0, while citrate inhibits binding in a complex, noncompetitive manner. In the presence of the ATP analog adenylyl imidodiphosphate, the enzyme-fructose 6-phosphate binding isotherm is sigmoidal at pH 7.0, but hyperbolic at pH 8.0. The characteristic sigmoidal initial velocity-fructose 6-phosphate isotherms for phosphofructokinase at pH 7.0, therefore, are due to an heterotropic interaction between ATP and fructose 6-phosphate binding sites which alters the homotropic interactions between fructose 6-phosphate binding sites. Thus the homotropic interactions between fructose 6-phosphate binding sites can give rise to positive, negative, or no cooperativity depending upon the pH, the aggregation state of the protein, and the metabolic effectors present. The available data suggest the regulation of phosphofructokinase involves a complex interplay between protein polymerization and homotropic and heterotropic interactions between ligand binding sites.  相似文献   

8.
Alkaline Bohr effect of human hemoglobin Ao   总被引:3,自引:0,他引:3  
  相似文献   

9.
An ultraviolet absorption difference spectrum that is typical of a change in ionization state (pKa 9.7 leads to greater than 11.5) of a tyrosyl residue has been observed on the binding between Streptomyces subtilisin inhibitor (SSI) and subtilisin BPN' [EC 3.4.21.14] at alkaline pH, ionic strength 0.1 M, at 25 degrees C (Inouye, K., Tonomura, B., and Hiromi, K., submitted). When the complex of SSI and subtilisin BPN' is formed at an ionic strength of 0.6 M and pH 9.70, the characteristic features of the protonation of a tyrosyl residue in the difference spectrum are diminished. These results suggest that the pKa-shift of a tyrosyl residue observed at alkaline pH and lower ionic strength results from an electrostatic interaction. Nitration of tyrosyl residues of SSI and of subtilisin BPN' was performed with tetranitromethane (TNM). By measurements of the difference spectra observed on the binding of the tyrosyl-residue-nitrated SSI and the native subtilisin BPN', and on the binding of the native SSI and the tyrosyl-residue-nitrated subtilisin BPN' and alkaline pH, the tyrosyl residue in question was shown to be one out of the five tyrosyl residues of pKa 9.7 of the enzyme. This tyrosyl residue was probably either Tyr 217 or Tyr 104 on the basis of the reactivities of tyrosyl residues of the enzyme with TNM and their locations on the enzyme molecule. Carboxyl groups of SSI were modified by covalently binding glycine methyl ester with the aid of water-soluble carbodiimide, in order to neutralize the negative charges on SSI. In the difference spectrum which was observed on the binding of subtilisin BPN' and the 5.3-carboxyl-group-modified SSI at alkaline pH, the characteristic features of the protonation of a tyrosyl residue were essentially lost, and the difference spectrum is rather similar to that observed on the binding of the native SSI and the enzyme at neutral pH. This phenomenon indicates that the pKa of a tyrosyl residue of the enzyme is shifted upwards by interaction with carboxyl group(s) of SSI on the formation of the enzyme-inhibitor complex.  相似文献   

10.
The relationship between fructose 2,6-bisphosphate (Fru-2,6-BP) activation and MgATP inhibition of rat liver phosphofructokinase has been comprehensively evaluated at pH 7. When either ligand is varied at a fixed concentration of the other, its influence on the concentration of fructose 6-phosphate (Fru-6-P) required to produce half-maximal velocity, Ka, is usually well described by the same simple, single-modifier linkage expression that described the actions of these ligands at pH 9. However, the effects of both ligands together cannot be described by the same overall linkage relationship that described their actions at pH 9. Specifically, despite an overall antagonistic relationship between the binding of MgATP and that of Fru-2,6-BP, very low concentrations of Fru-2,6-BP appear to facilitate the binding of MgATP to an appreciable degree. Also, MgATP at high concentration appears to inhibit the binding of Fru-2,6-BP to a significantly greater extent than its actions at lower concentration would predict. These additional features of MgATP-Fru-2,6-BP interaction have been incorporated into an overall linkage expression describing the actions of both MgATP and Fru-2,6-BP on Ka for Fru-6-P. The best fit parameters predict the data to within an average standard error of +/- 21%.  相似文献   

11.
We present dynamic light scattering (DLS) and hydrophobic dye-binding data in an effort to elucidate a molecular mechanism for the ability of gastric mucin to form a gel at low pH, which is crucial to the barrier function of gastric mucus. DLS measurements of dilute mucin solutions were not indicative of intermolecular association, yet there was a steady fall in the measured diffusion coefficient with decreasing pH, suggesting an apparent increase in size. Taken together with the observed rise in depolarized scattering ratio with decreasing pH, these results suggest that gastric mucin undergoes a conformational change from a random coil at pH >/= 4 to an anisotropic, extended conformation at pH < 4. The increased binding of mucin to hydrophobic fluorescent with decreasing pH indicates that the change to an extended conformation is accompanied by exposure of hydrophobic binding sites. In concentrated mucin solutions, the structure factor S(q, t) derived from DLS measurements changed from a stretched exponential decay at pH 7 to a power-law decay at pH 2, which is characteristic of a sol-gel transition. We propose that the conformational change facilitates cross-links among mucin macromolecules through hydrophobic interactions at low pH, which in turn leads to a sol-gel transition when the mucin solution is sufficiently concentrated.  相似文献   

12.
The enzyme rhodanese is greatly stabilized in the range pH 4-6, and samples at pH 5 are fully active after several days at 23 degrees C. This is very different from results at pH greater than 7, where there is significant loss of activity within 1 h. A pH-dependent conformational change occurs below pH 4 in a transition centered around pH 3.25 that leads slowly to inactive rhodanese at pH 3 (t 1/2 = 22 min at pH3). The inactive rhodanese can be reactivated by incubation under conditions required for detergent-assisted refolding of denatured rhodanese. The inactive enzyme at pH 3 has the maximum of its intrinsic fluorescence spectrum shifted to 345 nm from 335 nm, which is characteristic of native rhodanese at pH greater than 4. At pH 3, rhodanese shows increased exposure of organized hydrophobic surfaces as measured by 1,1'-bis(4-anilino)naphthalene-5,5'-disulfonic acid binding. The secondary structure is maintained over the entire pH range studied (pH 2-7). Fluorescence anisotropy measurements of the intrinsic fluorescence provide evidence suggesting that the pH transition produces a state that does not display greatly increased average flexibility at tryptophan residues. Pepsin digestibility of rhodanese follows the pH dependence of conformational changes reported by activity and physical methods. Rhodanese is resistant to proteolysis above pH 4 but becomes increasingly susceptible as the pH is lowered. The form of the enzyme at pH 3 is cleaved at discrete sites to produce a few large fragments. It appears that pepsin initially cleaves close to one end of the protein and then clips at additional sites to produce species of a size expected for the individual domains into which rhodanese is folded. Overall, it appears that in the pH range between pH 3 and 4, titration of groups on rhodanese leads to opening of the structure to produce a conformation resembling, but more rigid than, the molten globule state that is observed as an intermediate during reversible unfolding of rhodanese.  相似文献   

13.
K I Miller 《Biochemistry》1985,24(17):4582-4586
Oxygen binding by Octopus dofleini hemocyanin was examined under very nearly physiological conditions. The effects of pH, ionic composition, temperature, and aggregation were controlled so that the role each plays in modulating oxygen binding can be isolated. There is a very large effect of pH on affinity, the Bohr effect (delta log P50/delta pH = -1.7), which is the same at 10 and 20 degrees C. However, cooperativity is substantially altered over the same range of pHs at the two temperatures. The allosteric properties were examined by comparing the experimental data points to curves generated by use of the Monod-Wyman-Changeux model. A computer-fitting process was developed which allowed the individual allosteric parameters to be varied independently until the best fit could be determined. The relationship between kR and kT is responsible for the effect of pH on cooperativity. A change in the allosteric properties of the T form is primarily responsible for the differences due to temperature. Changing cation concentrations when the molecule is in the fully aggregated 51S form alters affinity without influencing cooperativity. The effect of Mg2+ is much greater than that of Na+. If the 51S decamer is dissociated to 11S monomers by removing divalent cations, oxygen binding is noncooperative. There is evidence for negative cooperativity, indicating heterogeneity of function within the subunit which contains seven oxygen binding domains. Association into decamers generates conformational change which results in a much wider range of allosteric function.  相似文献   

14.
Aryl sulfotransferase IV from rat liver has the very broad substrate range that is characteristic of the enzymes of detoxication. With the conventional assay substrates, 4-nitrophenol and PAPS, sulfation was considered optimal at pH 5.5 whereas the enzyme in the physiological pH range was curiously ineffective. These properties would seem to preclude a physiological function for this cytosolic enzyme. Partial oxidation of the enzyme, however, results not only in a substantial increase in the rate of sulfation of 4-nitrophenol at physiological pH but also in a shift of the pH optimum to this range and radically altered overall substrate specificity. The mechanism for this dependence on redox environment involves oxidation at Cys66, a process previously shown to occur by formation of a mixed disulfide with glutathione or by the formation of an internal disulfide with Cys232. Oxidation at Cys66 acts only as a molecular redox switch and is not directly part of the catalytic mechanism. Underlying the activation process is a change in the nature of the ternary complex formed between enzyme, phenol, and the reaction product, adenosine 3',5'-bisphosphate. The reduced enzyme gives rise to an inhibitory, dead-end ternary complex, the stability of which is dictated by the ionization of the specific phenol substrate. Ternary complex formation impedes the binding of PAPS that is necessary to initiate a further round of the reaction and is manifest as profound, substrate-dependent inhibition. In contrast, the ternary complex formed when the enzyme is in the partially oxidized state allows binding of PAPS and the unhindered completion of the reaction cycle.  相似文献   

15.
PHYSICAL STUDIES OF ISOLATED EUCARYOTIC NUCLEI   总被引:17,自引:8,他引:9       下载免费PDF全文
The degree of chromatin condensation in isolated rat liver nuclei and chicken erythrocyte nuclei was studied by phase-contrast microscopy as a function of solvent pH, K+ and Mg++ concentrations Data were represented as "phase" maps, and standard solvent conditions selected that reproducibly yield granular, slightly granular, and homogeneous nuclei Nuclei in these various states were examined by ultraviolet absorption and circular dichroism (CD) spectroscopy, low-angle X-ray diffraction, electron microscopy, and binding capacity for ethidium bromide Homogeneous nuclei exhibited absorption and CD spectra resembling those of isolated nucleohistone. Suspensions of granular nuclei showed marked turbidity and absorption flattening, and a characteristic blue-shift of a crossover wavelength in the CD spectra. In all solvent conditions studied, except pH < 2 3, low-angle X-ray reflections characteristic of the native, presumably superhelical, nucleohistone were observed from pellets of intact nuclei. Threads (100–200 A diameter) were present in the condensed and dispersed phases of nuclei fixed under the standard solvent conditions, and examined in the electron microscope after thin sectioning and staining Nuclei at neutral pH, with different degrees of chromatin condensation, exhibited similar binding capacities for ethidium bromide. These data suggest a model that views chromatin condensation as a close packing of superhelical nucleohistone threads but still permits condensed chromatin to respond rapidly to alterations in solvent environment.  相似文献   

16.
C-reactive protein (CRP) is a phylogenetically conserved protein; in humans, it is present in the plasma and at sites of inflammation. At physiological pH, native pentameric CRP exhibits calcium-dependent binding specificity for phosphocholine. In this study, we determined the binding specificities of CRP at acidic pH, a characteristic of inflammatory sites. We investigated the binding of fluid-phase CRP to six immobilized proteins: complement factor H, oxidized low-density lipoprotein, complement C3b, IgG, amyloid β, and BSA immobilized on microtiter plates. At pH 7.0, CRP did not bind to any of these proteins, but, at pH ranging from 5.2 to 4.6, CRP bound to all six proteins. Acidic pH did not monomerize CRP but modified the pentameric structure, as determined by gel filtration, 1-anilinonaphthalene-8-sulfonic acid-binding fluorescence, and phosphocholine-binding assays. Some modifications in CRP were reversible at pH 7.0, for example, the phosphocholine-binding activity of CRP, which was reduced at acidic pH, was restored after pH neutralization. For efficient binding of acidic pH-treated CRP to immobilized proteins, it was necessary that the immobilized proteins, except factor H, were also exposed to acidic pH. Because immobilization of proteins on microtiter plates and exposure of immobilized proteins to acidic pH alter the conformation of immobilized proteins, our findings suggest that conformationally altered proteins form a CRP-ligand in acidic environment, regardless of the identity of the protein. This ligand binding specificity of CRP in its acidic pH-induced pentameric state has implications for toxic conditions involving protein misfolding in acidic environments and favors the conservation of CRP throughout evolution.  相似文献   

17.
Dilatometric measurements of the volume changes accompanying the binding reactions of azide ion to human adult and pigeon methemoglobins as a function pH at 25°C demonstrate pH values of maximum volume change (pH ΔVmax) which are different for the different hemoglobins. pHΔVmax occurs at pH 6.7 for human methemoglobin A and at pH 7.7 for pigeon methemoglobin. The pHΔVmax occurs near the characteristic pH (pHch) of maximum enthalpy of the same binding reaction. It is shown that the large pH variation in ΔV can arise if the configuration of charged groups on the surface of the molecule is different in methemoglobin and methemoglobin complex. When such a difference in configuration exists the addition of the same number of protons to methemoglobin and methemoglobin complex will give rise to different changes in the partial molar volume of the two species.  相似文献   

18.
1. Complex formation at the general anion-binding site of the liver alcohol dehydrogenase subunit has been characterized by transient-state kinetic methods, using NADH as a reporter ligand. Equilibrium dissociation constants for anion binding at the site are reported. They conform basically to the lyotropic series of affinity order, with exceptionally tight binding of sulphate. The particular specificity for sulphate might be a general characteristic of anion-binding enzymic arginyl sites. 2. Anionic species of phosphate and pyrophosphate buffer solutions do not interact significantly with the general anion-binding site over the pH range 8-10. At lower pH, phosphate binding becomes significant due to complex formation with the monovalent H2PO4 species. The latter interaction corresponds to a dissociation constant of about 60 mM, indicating that phosphate binding is comparatively weak also at low pH. 3. It is concluded that previously reported pH dependence data for coenzyme binding to liver alcohol dehydrogenase cannot be much affected by coenzyme-competitive effects of buffer anion binding. Kinetic parameter estimates now determined for NADH binding in weakly buffered solutions agree within experimental precision with those obtained previously from measurements made in buffer solutions of 0.1 M ionic strength.  相似文献   

19.
Structural requirements for substrate binding to histidyl-tRNA synthetase from Salmonella typhimurium have been investigated using ATP analogues. Ki values and the relative binding affinity of the enzyme for these analogues have been determined in the tRNA aminoacylation reaction. The enzyme is highly specific for ATP: no binding was found for GTP, CTP, TTP and UTP. dATP is a very poor substrate for acylation of tRNA, with a Km 40-fold higher than that of ATP. Binding of adenosine 5'-triphosphate requires interactions of the amino group of adenosine and the sugar moiety; the 2' and the 5' positions of the ribose appear to be essential for recognition; the phosphate groups enhance the binding. AMP is a noncompetitive inhibitor with ATP. The interaction of histidyl-tRNA synthetase, a dimeric enzyme, with histidine and ATP was examined by fluorescence measurements at equilibrium and by equilibrium dialysis. Binding with L-histidine is significantly tighter at pH 6 than at pH 7, while the ATP binding is independent of pH. The stoichiometry was measured at pH 6 than at pH 7, while the ATP binding is independent of pH. The stoichiometry was measured at pH 7.5 by equilibrium dialysis and is 1 mol ATP/mol enzyme and, variably, close to 2 or 1 mol histidine/mol enzyme.  相似文献   

20.
In the absence of experimental methods for determining concentrations of species in protein-ligand binding, it is not possible to determine the thermodynamic properties of species directly. However, this article on a simple reaction system shows that measurements of the average number of oxygen molecules bound at various T, pH and concentrations of molecular oxygen can be used to calculate thermodynamic properties of species. The simple system considered has some of the characteristics of the binding of oxygen by hemoglobin, but it has been simplified so that the method for obtaining thermodynamic information can be clarified. A table of standard thermodynamic properties of species is the most efficient way to store thermodynamic information on a reaction system. All the standard further transformed thermodynamic properties at specified T, pH and concentrations of molecular oxygen, all the standard transformed thermodynamic properties at specified T and pH, and all the standard thermodynamic properties of species at a specified temperature can be calculated. These calculations are based on the fact that the mathematical function for the standard further transformed Gibbs energy of the system contains all the thermodynamic information on the system. These properties are all interrelated by Maxwell equations.  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