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.
The rates of hydrolysis of TrEE, TEE, and ATEE1 by α-chymotrypsin were determined in media of variable dielectric strength. Many substances which modify the dielectric constant of the medium, exert additional specific effects on the reaction rate, noticeable at more or less elevated concentrations. Notwithstanding, it is possible to differentiate the dielectric and specific effects by comparing the rates in solvents of distinct nature at relatively low concentrations. Thus, the effect of varying the dielectric strength could be studied within wider ranges (ΔD = 20 with TrEE and ca. 28 with ATEE) than in the previous study of trypsin (ΔD = 12). The dielectric effect on α-chymotrypsin is the opposite of that observed with trypsin. In both cases there is a linear relationship between the logarithm of the rate of hydrolysis and the reciprocal of the dielectric constant. The slope is negative with α-chymotrypsin and positive with trypsin. According to expressions relating the dielectric constant to the rate in non-enzymatic reactions, the behavior of α-chymotrypsin is like that of a negative ion, while trypsin behaves as a positive ion. The enzyme activity appears to depend upon the arrangement of charges in the enzyme and substrate molecules, rather than on the presence of certain atomic groupings in the substrate.  相似文献   

2.
There is a mobile equilibrium between the native and denatured forms of trypsin which depends on the concentrations of acid, alkali, and alcohol and on the temperature. The heat of denaturation in 0.01 N hydrochloric acid calculated from the effect of temperature on the equilibrium constant is –67,600 calories per mole.  相似文献   

3.
The contribution of electrostriction of the solvent to the stabilization of the negatively charged tetrahedral transition state of a trypsin-catalyzed reaction was probed by means of kinetic studies involving high-pressure and solvent dielectric constant. A good correlation was observed between the increased catalytic efficiency of trypsin and the decreased solvent dielectric constant. When the dielectric constant of the solvents was lowered by 4.68 units, the loss of activation energy and that of free energy of activation were 2.26 kJ/mol and 3.09 kJ/mol, respectively. The activation volume for kcat decreased significantly as the dielectric constant of the solvent decreased, indicating that the degree of electrostriction of the solvent around the charged tetrahedral transition state has been enhanced. These observations demonstrate that the increase in the catalytic efficiency of the trypsin reaction with decreasing dielectric constant resulted from the stabilization of electrostatic energy for the formation of an oxyanion hole, and this stabilization was caused by the increase of electrostricted water around the charged tetrahedral transition state. Therefore, we conclude that control of the solvent dielectric constant can stabilize the tetrahedral transition state, and this lowers the activation energy.  相似文献   

4.
Qualitative differences in the active center of rat trypsins 1 and 2 resulted in different ratios of Kcat for N1-tosyl-l-arginine methyl ester vs Kcat for N1-benzoyl-l-arginine ethyl ester. These ratios were 2.5 for trypsin 1 and 1.2 for trypsin 2.Substrate activation with N1-tosyl-l-arginine methyl ester enhanced the catalytic rate constant of rat trypsin 1 2.5-fold and that of rat trypsin 2 only 1.5-fold. The increase in the catalytic rate constant found with N1-benzoyl-l-arginine ethyl ester was the same (1.5-fold) for both trypsins. Consequently, at 20 mm substrate concentration, trypsin 1 catalyzed the esterolysis of N1-tosyl-l-arginine methyl ester 4.5 times faster than that of N1-benzoyl-l-arginine ethyl ester, while trypsin 2 was only 1.3 times more efficient with the first substrate.Furthermore, the activation of both rat enzymes by N-acetyl-l-tyrosine ethyl ester was even more effective than that obtained with the two cationic esters; the maximum rates of hydrolysis of this neutral substrate by trypsins 1 and 2 were enhanced 120- and 50-fold, respectively, by high concentrations of N-acetyl-l-tyrosine ethyl ester.  相似文献   

5.
(1) Using the pulse-radiolysis and stopped-flow techniques, the reactions of iron-free (porphyrin) cytochrome c and native cytochrome c with cytochrome aa3 were investigated. The porphyrin cytochrome c anion radical (generated by reduction of porphyrin cytochrome c by the hydrated electron) can transfer its electron to cytochrome aa3. The bimolecular rate constant for this reaction is 2·107 M?1·s?1 (5 mM potassium phosphate, 0.5% Tween 20, pH 7.0, 20°C). (2) The ionic strength dependence of the cytochrome c-cytochromeaa3 interaction was measured in the ionic strength range between 40 and 120 mM. At ionic strengths below 30 mM, a cytochrome c-cytochrome aa3 complex is formed in which cytochrome c is no longer reducible by the hydrated electron. A method is described by which the contributions of electrostatic forces to the reaction rate can be determined. (3) Using the stopped-flow technique, the effect of the dielectric constant (?) of the reaction medium on the reaction of cytochrome c with cytochrome aa3 was investigated. With increasing ? the second-order rate constant decreased.  相似文献   

6.
1. A study has been made of the equilibrium existing between trypsin and the substances formed in the digestion of proteins which inhibit its action. 2. This substance could not be obtained by the hydrolysis of the proteins by acid or alkali. It is dialyzable. 3. The equilibrium between this substance (inhibitor) and trypsin is found to agree with the equation, trypsin + inhibitor ⇌ trypsin-inhibitor The equilibrium is reached instantaneously and is independent of the substrate concentration. If it be further assumed that the rate of hydrolysis is proportional to the concentration of the free trypsin and that the equilibrium conforms to the law of mass action, it is possible to calculate the experimental results by the application of the law of mass action. 4. The equilibrium has been studied by varying (a) the concentration of the inhibiting substance, (b) the concentration of trypsin, (c) the concentration of gelatin, and (d) the concentration of trypsin and inhibitor (the relative concentration of the two remaining the same). In all cases the results agree quantitatively with those predicted by the law of mass action. 5. It was found that the percentage retarding effect of the inhibiting substance on the rate of hydrolysis is independent of the hydrogen ion concentration between pH 6.3 and 10.0. 6. The fact that the experimental results agree with the mechanism outlined under 3, is contrary to the assumption that any appreciable amount of trypsin is combined with the gelatin at any one time; i.e., the velocity of the hydrolysis must depend on the time required for such a compound to form rather than for it to decompose. 7. The experiments may be considered as experimental proof of the validity of Arrhenius'' explanation of Schütz''s rule as applied to trypsin digestion. 8. Inactivated trypsin does not enter into the equilibrium.  相似文献   

7.
The kinetics of the reaction between cytochrome c peroxidase and fluoride was investigated as a function of ionic strength over the pH range 4 to 8. The ionic strength was varied between 0.01 and 0.10 m. At 0.01 m ionic strength, the reaction rates were determined between pH 2.7 and 9.2. A consideration of the ionic strength and pH dependence of the association rate constant for the fluoride-cytochrome c peroxidase reaction leads to the conclusion that hydrofluoric acid is the only significant reactive form of the ligand between pH 2.5 and 8. Above pH 8, binding of fluoride anion contributes to the apparent association rate even though the pH-independent rate constant for fluoride anion binding is more than 3 × 105 times smaller than the rate constant for hydrofluoric acid binding.  相似文献   

8.
The kinetics of the hydrogen-deuterium exchange reaction in a trypsin inhibitor (Kunitz) from soya bean have been followed by infrared absorption measurements in aqueous solutions at various temperatures and pH values. It was found that, in every case, 49% of the total peptide hydrogen atoms exchange relatively slowly. This amount corresponds to 83 peptide groups per molecule, and this is considered to be equal to the number of peptide NH groups involved in hydrogen bonds with the carbonyls of other peptide groups in the protein molecule in its native form. Each rate constant (k) determined at pH 2.75 for this category of the NH groups is in good agreement with the value expected from an idea that the breaking of the peptide-peptide hydrogen bonds takes place very slowly, and that this is the rate-determining process in the hydrogen-deuterium exchange reaction. Thus, by ultraviolet absorption measurements at 297 nm, the equilibrium constant of the native and denatured forms has been determined in the temperature range from 42 to 53.5 °C, as well as the reaction rate of reaching equilibrium from an off-equilibrium state. From these data the rate constant (k1) of the denaturation reaction is determined, and the k1 value is found to be practically equal to the hydrogen exchange rate constant (k). The Arrhenius plot of this rate constant (k) gives a straight line in the 25 to 55 °C region, and this gives a value of 48.6 kcal/mol for the activation energy of the denaturation reaction. The rate of this reaction is found to be very low at 25 °C; its half-life is about eleven days. Infrared absorption spectra observed in the amide I region suggest that the very slow denaturation of this protein is accompanied by a conformation change from an α-helix to a β-form. The number of the peptide groups involved in this αβ change is estimated to be 9 ± 3.  相似文献   

9.
This investigation concerns the effect of certain physical factors—viscosity, dielectric constant, ionic strength, and temperature of the medium—on the reaction of hydrogen peroxide and ferrocytochrome c in the presence of the enzyme horse-radish peroxidase. From study of the effects of viscosity and dielectric constant, it was concluded that the reaction between the secondary complex of hydrogen peroxide and enzyme on the one hand and ferrocytochrome c on the other is controlled by diffusion in media of high viscosity and by electrostatic effects at low viscosities. With respect to ionic strength, the data at pH 4.7 indicated a dipole-dipole interreaction. The temperature dependence of the over-all reaction had a Q10 of 1.25.  相似文献   

10.
L R Holman  D O Jordan 《Biopolymers》1972,11(8):1661-1684
DNA which has been heat denatured in the presence of Cu++ ions can be completely and rapidly renatured by increasing the ionic strength of the solution above a critical value. A kinetic study of this renaturation recation was carried out by following the associated UV absorbance change and also by following the change in free Cu++ ion concentration by means of a specific Cu++ ion activity electrode. The data obtained could be fitted to first-order kinetics for a considerable extent of the reaction and the rate constant was found to increase with temperature and ionic strength, but to decrease markedly as the bulk viscosity of the solution was increased. At temperatures greater than 5°C the reaction rate depended on the time elapsing between denaturation and the commencement of the renaturation reaction. As there was good agreement between the rate constants obtained by following the decrease in hyperchromism and by following the increase in free Cu++ ion concentration, it is concluded that under the conditions employed, the rate of renaturation is determined by the rate of release of Cu++ ions from the denatured DNA-Cu++ complex.  相似文献   

11.
The hydrolysis reaction of N α-benzoyl-L-arginine ethyl ester catalyzed by trypsin from pig pancreas was comparatively studied in an aqueous buffer solution and in the system of reversed micelles of Aerosol OT in octane (pH 8.5) to determine the mechanisms of influence of the enzyme microenvironment on the rate constants of the elementary stages of the enzymatic reaction. The temperature dependences of the catalytic constant k cat and the rate constant of the second order k cat/K m (s, catalysis efficiency) allowed the determination of the rate constants and the activation energy of elementary stages of the enzymatic reaction. It was revealed that a decrease in the efficiency of catalytic action of trypsin in reverse micelles in comparison with an aqueous solution is first of all determined by a decrease in the rate constant of formation of the enzyme-substrate complex k 1. Possible mechanisms of the effect of the microenvironment on the elementary stages of catalytic action of the enzyme are discussed.  相似文献   

12.
The kinetics of electron transfer between the isolated enzymes of cytochrome c1 and cytochrome c have been investigated using the stopped-flow technique. The reaction between ferrocytochrome c1 and ferricytochrome c is fast; the second-order rate constant (k1) is 3.0 · 107 M?1 · s?1 at low ionic strength (I = 223 mM, 10°C). The value of this rate constant decreases to 1.8 · 105 M?1 · s?1 upon increasing the ionic strength to 1.13 M. The ionic strength dependence of the electron transfer between cytochrome c1 and cytochrome c implies the involvement of electrostatic interactions in the reaction between both cytochromes. In addition to a general influence of ionic strength, specific anion effects are found for phosphate, chloride and morpholinosulphonate. These anions appear to inhibit the reaction between cytochrome c1 and cytochrome c by binding of these anions to the cytochrome c molecule. Such a phenomenon is not observed for cacodylate. At an ionic strength of 1.02 M, the second-order rate constants for the reaction between ferrocytochrome c1 and ferricytochrome c and the reverse reaction are k1 = 2.4 · 105 M?1 · s?1 and k?1 = 3.3 · 105 M?1 · s?1, respectively (450 mM potassium phosphate, pH 7.0, 1% Tween 20, 10°C). The ‘equilibrium’ constant calculated from the rate constants (0.73) is equal to the constant determined from equilibrium studies. Moreover, it is shown that at this ionic strength, the concentrations of intermediary complexes are very low and that the value of the equilibrium constant is independent of ionic strength. These data can be fitted into the following simple reaction scheme: cytochrome c2+1 + cytochrome c3+ai cytochrome c3+1 + cytochrome c2+.  相似文献   

13.
Reaction of Oxyhemoglobin with Carbon Monoxide   总被引:1,自引:0,他引:1       下载免费PDF全文
The reaction of oxyhemoglobin and carbon monoxide was studied kinetically at pH 7.8 in a variety of suspending media. The dielectric constant of the suspending media, as well as the viscosity (and hence the Fick diffusion coefficients), was varied with the use of glycine, glycerol, and sucrose. The results showed that the reaction was unaltered by the various additions to the media, provided that the pO2 and the concentration of carbon monoxide were held constant. Since the concentration of oxygen varies from medium to medium at constant pO2 while the pCO varies at constant concentration of carbon monoxide, the differences in the reactions with oxygen and carbon monoxide were emphasized. The lack of variation of the rate constants with changes in dielectric constant can be interpreted as indicating that electrostatic effects are unimportant in this reaction.  相似文献   

14.
A study was made on the effect of glycine on systems involving trypsin and BAEE1 or TSAME on the one hand, or α-chymotrypsin with any of the substrates BAEE, TEE, or PEE, on the other. In all cases there was a linear relationship between the rate logarithm and the reciprocal of the dielectric constant of the glycine solution. The slopes were positive in the reactions of trypsin. In those catalyzed by α-chymotrypsin, the slopes were positive at pH 6.5 or lower, and negative at pH 7.5. However, the effects of glycine differ quantitatively from those of urea or other solvents. The presence of salt modifies somewhat the glycine effects. A low ionic strength increases the effect of glycine on trypsin, but if the inhibition caused by the ionic strength is relatively strong, the addition of glycine partially neutralizes the salt effect. Addition of salt to systems containing α-chymotrypsin always resulted in a diminished effect of glycine. An attempt is made to interpret the anomalies of glycine effects on the basis of its dipolar ion structure.  相似文献   

15.
The binding of pancreatic ribonuclease-A by denatured DNA, native DNA, poly-dA, and poly-dT, has been studied by a gel filtration method. With denatured DNA at pH 7.5, ionic strength 0.053M, there is one binding site per 12 nucleotides and the equilibrium binding constant per site is 9.7 × 104 l./mole. The binding constant increases by a factor of 8 as the pH is decreased from 8 to 7. The strength of the binding of denatured DNA increases with decreasing ionic strength. At pH 7.5, native DNA binds about ? as strongly as does denatured DNA. The binding affinity increases in the order poly-dA, denatured DNA, and poly-dT. These results support the view that the binding of denatured DNA involves both electrostatic interactions between the negatively charged polynucleotide and the positively charged protein, and an interaction of the protein with a pyrimidine residue of the denatured DNA, and thus that the binding is basically similar to that between RNAse and its substrate RNA.  相似文献   

16.
1. An equation is derived for the calculation of a constant which, experimental results indicate, may be a more reliable index of the antiproteolytic activity of serum than those equations hitherto used. 2. (a) Intramuscular administration of trypsin resulted in a slow rise in the antiproteolytic activity of the serum, followed by a lesser decline. (b) Intravenous administration resulted in no appreciable variation. (c) Oral administration resulted in a rapid rise, which was sustained during the period of administration. (d) Intramuscular, intravenous, or oral administration of denatured trypsin resulted in no appreciable variation. (e) The extent of the local necrosis following subcutaneous injection of trypsin varied inversely with the antiproteolytic activity of the serum. 3. The experimental results indicate that the products of protein hydrolysis in the intestine and parenterally are an important factor in the antiproteolytic activity of the serum. They also indicate that antibodies to trypsin are not an important factor in the antiproteolytic activity of the serum.  相似文献   

17.
Intracellular free calcium concentration in the sea urchin egg was calculated to increase from 0.1 mM in an unfertilized egg to 1 mM in a fertilized egg 10 min after fertilization, based on measurement of the dissociation constant between free calcium and sea urchin egg homogenate. The dissociation constant between free calcium (dialyzable calcium) and homogenate of sea urchin eggs was measured by means of dialysis equilibrium. The dissociation constant of the unfertilized egg was about 10–4 M and that of the fertilized egg was about 10–3 M in three species of sea urchin, Hemicentrotus pulcherrimus, Anthocidaris crassispina, and Pseudocentrotus depressus. An increase in the dissociation constant of the unfertilized egg homogenate was observed after the addition of calcium ion at a concentration above 0.3 mM, the dissociation constant becoming the same as that observed in the fertilized egg homogenate after the administration of CaCl2 at a concentration above 1 mM. Sodium ion also caused a decrease in the calcium-binding ability of the unfertilized egg homogenate. Therefore, penetration of calcium ion or sodium ion upon fertilization might induce an increase in the dissociation constant and then intracellular concentration of free calcium would increase at fertilization. Almost all calcium-binding ability of the egg homogenate was found in the microsomal fraction, and the substance which bound calcium was thought to be protein in nature, since trypsin could decrease the level of calcium-binding substance in the homogenate of the eggs.  相似文献   

18.
Normal and transformed baby hamster kidney (BHK) cells attach to Falcon polystyrene with the same first order rate constant. The longer the cells are attached to the bottles, the more difficult they are to remove. Sulfhydryl (—SH) binding reagents inhibit both the attachment of BHK cells and the increase in adhesive strength of attached cells. Attached BHK cells bind fewer molecules of [1-14C]N-ethylamleimide (an —SH binding reagent) than do suspended cells. Incubation of cells with high concentrations of trypsin results in a reversible loss of cell adhesiveness. The recovery of adhesiveness of trypsin-treated cells is inhibited by cycloheximide.  相似文献   

19.
The effects of K+, Na+ and ATP on the gastric (H+ + K+)-ATPase were investigated at various pH. The enzyme was phosphorylated by ATP with a pseudo-first-order rate constant of 3650 min?1 at pH 7.4. This rate constant increased to a maximal value of about 7900 min?1 when pH was decreased to 6.0. Alkalinization decreased the rate constant. At pH 8.0 it was 1290 min?1. Additions of 5 mM K+ or Na+, did not change the rate constant at acidic pH, while at neutral or alkaline pH a decrease was observed. Dephosphorylation of phosphoenzyme in lyophilized vesicles was dependent on K+, but not on Na+. Alkaline pH increased the rate of dephosphorylation. K+ stimulated the ATPase and p-nitrophenylphosphatase activities. At high concentrations K+ was inhibitory. Below pH 7.0 Na+ had little or no effect on the ATPase and p-nitrophenylphosphatase, while at alkaline pH, Na+ inhibited both activities. The effect of extravesicular pH on transport of H+ was investigated. At pH 6.5 the apparent Km for ATP was 2.7 μM and increased little when K+ was added extravesicularly. At pH 7.5, millimolar concentrations of K+ increased the apparent Km for ATP. Extravesicular K+ and Na+ inhibited the transport of H+. The inhibition was strongest at alkaline pH and only slight at neutral or acidic pH, suggesting a competition between the alkali metal ions and hydrogen ions at a common binding site on the cytoplasmic side of the membrane. Two H+-producing reactions as possible candidates as physiological regulators of (H+ + K+)-ATPase were investigated. Firstly, the hydrolysis of ATP per se, and secondly, the hydration of CO2 and the subsequent formation of H+ and HCO3?. The amount of hydrogen ions formed in the ATPase reaction was highest at alkaline pH. The H+/ATP ratio was about 1 at pH 8.0. When CO2 was added to the reaction medium there was no change in the rate of hydrogen ion transport at pH 7.0, but at pH 8.0 the rate increased 4-times upon the addition of 0.4 mM CO2. The results indicate a possible co-operation in the production of acid between the H+ + K+-ATPase and a carbonic anhydrase associated with the vesicular membrane.  相似文献   

20.
The reaction of triethanolamine (TEA) with active substrates—p-nitrophenyl esters and cinnamoyl imidazole (CI)—is catalyzed by divalent heavy metal ions. With Hg2+, rate enhancements of 100–1000 (depending on the substrate) were observed, the overall rate constants of substrate decomposition thus exceeding those of spontaneous hydrolysis up to 100,000-fold. The predominant active species at low L:M ratio was found to be the Hg-(TEA)2 complex. The dependence of the reaction rate upon excess of amino alcohol—at constant Hg2+ concentration—is attributable to formation of another active complex—Hg-(TEA)3.The high reactivity of the system is due to the alcoholate group of metal-bound TEA, whose pK has been lowered by the proximity of the metal ion. This labile nucleophilic alcoholate attacks the substrate causing its alcoholysis and forming O-acyl-TEA. The lability of the metal-alcoholate bond can be enhanced by low concentrations of halide ions, thus causing up to 5-fold additional increase in alcoholysis rate. Higher halide ion concentrations cause inhibition, probably due to formation of inactive HgX2 molecules.Presumably an important role of the metal ion in metalloenzymes is to affect the decrease in the pK value of a reactive group so that it can exhibit activity under physiological conditions.  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