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.
Violamycin BI is an anthracycline derivative with two sugars hanging on, each of them carries one positive charge. It dimerizes under conditions, which depend on the concentration of the antibiotic, pH and the ionic strength of the solution. By keeping a constant pH in a phosphate-EDTA buffer, the rate constants of violamycin BI dimerization were determined at various ionic strengths by temperature jump method. The dimerization constant Kd, resulting from the ratio of these rate constants, confirmed the values obtained spectrophotometrically in this study or elsewhere. The influence of ionic strength (0.02-0.2 M) on the rate constant values suggested to us some speculations on the reaction mechanism of the dimerization, in which, the specific mutual orientation of the monomers in the encounter, and perhaps a specific conformation of their side groups is required before a stabilizing action of the binding forces sets in.  相似文献   

2.
Both the forward and backward reactions of xylose isomerase (Sweetzyme Q) with xylose and glucose as substrates have been studied in terms of kinetics and thermodynamics. The relationship between the two reactions can thus be determined. Much attention has been given to the reaction with xylose as substrate. The optimal conditions of the xylose reaction in terms of pH, buffer, metal ions, substrate concentration, temperature, and ionic strength have been determined. These findings did not differ much from those reported for the glucose reaction. Equilibrium constants for the aldose to ketose conversion were more favorable in the case of glucose. The results obtained with continuous isomerization of xylose in columns packed with either Sweetzyme Q or Taka-Sweet were very similar to those obtained from batch isomerization processes. Particle size had a definite effect on reaction rate, which indicates that diffusion limitations do occur with the immobilized enzyme particles. Heat stability of Sweetzyme Q was good with t(1/2) of 118, 248, and 1200 h at 70, 55, and 40 degrees C, respectively. A novel method for the separation of xylose-xylulose mixtures with water as eluant on a specially prepared Dowex 1 x 8 column was developed. This technique has the capability of producing pure xylulose for industrial or research applications. A writ for a patent regarding this technique is at present prepared.  相似文献   

3.
K L Kim  D S Kang  L B Vitello  J E Erman 《Biochemistry》1990,29(39):9150-9159
The steady-state kinetics of the cytochrome c peroxidase catalyzed oxidation of horse heart ferrocytochrome c by hydrogen peroxide have been studied at both pH 7.0 and pH 7.5 as a function of ionic strength. Plots of the initial velocity versus hydrogen peroxide concentration at fixed cytochrome c are hyperbolic. The limiting slope at low hydrogen peroxide give apparent bimolecular rate constants for the cytochrome c peroxidase-hydrogen peroxide reaction identical with those determined directly by stopped-flow techniques. Plots of the initial velocity versus cytochrome c concentration at saturating hydrogen peroxide (200 microM) are nonhyperbolic. The rate expression requires squared terms in cytochrome c concentration. The maximum turnover rate of the enzyme is independent of ionic strength, with values of 470 +/- 50 s-1 and 290 +/- 30 s-1 at pH 7.0 and 7.5, respectively. The limiting slope of velocity versus cytochrome c concentration plots provides a lower limit for the association rate constant between cytochrome c and the oxidized intermediates of cytochrome c peroxidase. The limiting slope varies from 10(6) M-1 s-1 at 300 mM ionic strength to 10(8) M-1 s-1 at 20 mM ionic strength and extrapolates to 5 x 10(8) M-1 s-1 at zero ionic strength. The data are discussed in terms of both a two-binding-site mechanism and a single-binding-site, multiple-pathway mechanism.  相似文献   

4.
S Hahm  B Durham  F Millett 《Biochemistry》1992,31(13):3472-3477
The reactions of yeast cytochrome c peroxidase with horse cytochrome c derivatives labeled at specific lysine amino groups with (dicarboxybipyridine)(bisbipyridine)ruthenium(II) [Ru(II)] were studied by flash photolysis. All of the derivatives formed complexes with cytochrome c peroxidase compound I (CMPI) at low ionic strength (2 mM sodium phosphate, pH 7). Excitation of Ru(II) to Ru(II*) with a short laser flash resulted in electron transfer to the ferric heme group in cytochrome c, followed by electron transfer to the radical site in CMPI. This reaction was biphasic and the rate constants were independent of CMPI concentration, indicating that both phases represented intracomplex electron transfer from the cytochrome c heme to the radical site in CMPI. The rate constants of the fast phase were 5200, 19,000, 55,000, and 14,300 s-1 for the derivatives modified at lysines 13, 25, 27, and 72, respectively. The rate constants of the slow phase were 260, 520, 200, and 350 s-1 for the same derivatives. These results suggest that there are two binding orientations for cytochrome c on CMPI. The binding orientation responsible for the fast phase involves a geometry that supports rapid electron transfer, while that for the slow phase allows only slow electron transfer. Increasing the ionic strength up to 40 mM increased the rate constant of the slow phase and decreased that of the fast phase. A single intracomplex electron transfer phase with a rate constant of 2800 s-1 was observed for the lysine 72 derivative at this ionic strength. When a series of light flashes was used to titrate CMPI to CMPII, the reaction between the cytochrome c derivative and the Fe(IV) site in CMPII was observed. The rate constants for this reaction were 110, 250, 350, and 140 s-1 for the above derivatives measured in low ionic strength buffer.  相似文献   

5.
The cytochrome P450 (CYP) reaction mechanism often yields a broad array of coupled and uncoupled products from a single substrate. While it is well known that reaction conditions can drastically affect the rate of P450 catalysis, their effects on regioselectivity and coupling are not well characterized. To investigate such effects, the CYP1A2 oxidation of 7-ethoxymethoxy-3-cyanocoumarin (EOMCC) was examined as a function of buffer type, buffer concentration, pH, and temperature. A high-throughput, optical method was developed to simultaneously measure the rate of substrate depletion, NADPH depletion, and generation of the O-dealkylated product. Increasing the phosphate buffer concentration and temperature increased both the NADPH and EOMCC depletion rates by 6-fold, whereas coupling was constant at 7.9% and the regioselectivity of O-dealkylation to other coupled pathways was constant at 21.7%. Varying the buffer type and pH increased NADPH depletion by 2.5-fold and EOMCC depletion by 3.5-fold; however, neither coupling nor regioselectivity was constant, with variations of 14.4% and 21.6%, respectively. Because the enzyme–substrate binding interaction is a primary determinant of both coupling and regioselectivity, it is reasonable to conclude that ionic strength, as varied by the buffer concentration, and temperature alter the rate without affecting binding while buffer type and pH alter both.  相似文献   

6.
R L Blakley  L Cocco 《Biochemistry》1985,24(18):4772-4777
Stopped-flow measurements of protein fluorescence quenching when methotrexate (MTX) binds to dihydrofolate reductase (isoenzyme II) of Streptococcus faecium (SFDHFR II) analyze as the sum of two differentials: a rapid binding phase and a second phase for which the observed rate constant is independent of methotrexate concentration. Analysis of variation of the ratio of the amplitude of the fast and slow phases with methotrexate concentration indicates that the second phase is an isomerization of the initial binary complex. At pH 7.3, the equilibrium constant for this isomerization is 21.9, and the forward and reverse rate constants are 0.57 and 0.026 s-1, respectively. Similar results were obtained for binding of 3-deazamethotrexate to SFDHFR II, but the forward rate constant is greater (2.9 s-1 at pH 7.3). The equilibrium constants for these isomerizations are pH independent, but the rate constants decrease as the pH is raised, probably due to deprotonation of one or more groups on the enzyme. Analysis of progress curves obtained by the development of inhibition when SFDHFR II is added last to reaction mixtures containing dihydrofolate, NADPH, and MTX gives an association constant for initial reactions of 4.3 X 10(7) M-1. Since a preliminary estimate of the association constant for the binding reaction is 7.6 X 10(5) M-1, this suggests an isomerization of the ternary complex(es) with an equilibrium constant of about 56. In addition, analysis of the progress of development of inhibition indicates a further very slow isomerization with equilibrium constant 419 and forward rate constant 2.6 min-1.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

7.
S Loo  J E Erman 《Biochemistry》1975,14(15):3467-3470
The rate of the reaction between cytochrome c peroxidase and hydrogen peroxide was investigated using the stopped-flow technique. The apparent bimolecular rate constant was determined between pH 3.3 and pH 11 as a function of ionic strength. The pH dependence of the apparent bimolecular rate constant can be explained by assuming that two ionizable groups on the enzyme strongly influence the rate of the reaction. At 0.1 M ionic strength, a group with a pKa of 5.5 must be unprotonated and a group with a pKa of 9.8 must be protonated for the enzyme to react rapidly with hydrogen peroxide. The apparent acid dissociation constants depend upon the ionic strength. The true bimolecular rate constant has a value of (4.5 +/- 0.3) X 10(7) M-1 sec-1 and is independent of ionic strength.  相似文献   

8.
Triblock copolymers of functionalized poly(lactic acid)-b-poly(ethylene glycol)-b-poly(lactic acid) (PLA-b-PEG-b-PLA) have been widely investigated as precursors for fabricating resorbable polymeric drug delivery vehicles and tissue engineering scaffolds. Previous studies show degradation and erosion behavior of PLA-b-PEG-b-PLA hydrogels to rely on macromer chemistry as well as structural characteristics of the cross-linked networks. In this research, the degradation kinetics of diacrylated PLA-b-PEG-b-PLA copolymers as soluble macromers and cross-linked gels are directly compared as a function of macromer concentration, buffer pH, and ionic strength. The pseudo first-order rate constants for degradation of soluble macromers increase with water concentration and show a minimum at intermediate pH values, but are insensitive to ionic strength. The degradation rate constants for covalently cross-linked gels display a greater sensitivity to local water concentration and a minimum at lower pH values than corresponding soluble macromers. In addition, ionic strength significantly affects the rate of gel degradation due to the direct correlation between the degree of network ionization and gel water content.  相似文献   

9.
The effects of pH and ionic strength on the equilibrium constants and rate constants (binding and dissociation rate constants) between riboflavin binding protein (RBP) and flavins (riboflavin, 3-carboxymethylriboflavin [CMRF], and FMN) were studied by fluorometry. The equilibrium constant and the binding rate constant between RBP and riboflavin were pH-independent between pH 6 and 9, and both constants were also independent of the ionic strength, while the constants between RBP and CMRF or FMN were dependent on both pH and ionic strength. The dissociation rate constants between RBP and the flavins used here were not so dependent on pH and ionic strength in the pH region 6 to 9, and the patterns of pH profiles as a whole were similar to each other, although the constants for FMN were about 30-60 times larger than those for CMRF or riboflavin. RBP had lower affinity for FMN than for riboflavin in the neutral pH region, which is based on the small binding rate constant and the large dissociation rate constant for FMN. The former is due to an electrostatic repulsion force between negative net charges of RBP and the phosphate group of FMN, and the latter is due to steric interference by the phosphate group of FMN.  相似文献   

10.
11.
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   

12.
The parameters in steady-state or rapid-equilibrium rate equations for enzyme-catalyzed reactions depend on the temperature, pH, and ionic strength, and may depend on the concentrations of specific species in the buffer. When the complete rate equation (i.e. the equation with parameters for the reverse reaction as well as the forward reaction) is determined, there are one or more Haldane relations between some of the kinetic parameters and the apparent equilibrium constant for the reaction that is catalyzed. When the apparent equilibrium constant can be calculated from the kinetic parameters, the equilibrium composition can be calculated. This is remarkable because the kinetic parameters all depend on the properties of the enzymatic site, but the apparent equilibrium constant and the equilibrium composition do not. The effects of ionic strength and pH on the unoccupied enzymatic site and the occupied enzymatic site have to cancel in the Haldane relation or in the calculation of the apparent equilibrium constant using the rate constants for the steps in the mechanism. Several simple enzymatic mechanisms and their complete rate equations are discussed.  相似文献   

13.
The rate of hydrolysis of the imido ester methyl acetimidate and its rate of amidination of denatured aldolase were investigated under different conditions of temperature, pH and ionic strength. Both rate constants increase greatly with temperature, whereas ionic strength has no effect on either. The effect of pH is more complex. Between pH 6.8 and 8.8 the rate of hydrolysis decreases and the rate of amidination increases. These results are discussed in terms of the reaction mechanisms involved.  相似文献   

14.
The mechanism of binding of two antagonists, 3-quinuclidinyl benzilate and N-methyl-4-piperidinyl benzilate, to the muscarinic receptor was studied. The pseudo-first order rate constant of association showed a hyperbolic dependence on the concentration of the antagonist(s) indicating that the interaction involves two equilibria. The first binding equilibrium is reached rapidly and is characterized by dissociation constants 2.7 +/- 0.4 nM and 6.7 +/- 2.5 nM in phosphate buffer (0.05 M, pH = 7.4) for 3-quinuclidinyl benzilate and N-methyl-4-piperidinyl benzilate, respectively. The first binding equilibrium is followed by a slower isomerization step of the receptor . antagonist complex. The equilibrium constants for the isomerization step of the complex for both ligands were about 0.15. The overall constant of binding obtained as the product of the above constants shows good agreement with the results of equilibrium binding studies.  相似文献   

15.
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   

16.
Kinetics of the daunomycin--DNA interaction   总被引:2,自引:0,他引:2  
The kinetics of the interaction of daunomycin with calf thymus DNA are described. Stopped-flow and temperature-jump relaxation methods, using absorption detection, were used to study the binding reaction. Three relaxation times were observed, all of which are concentration dependent, although the two slower relaxations approach constant values at high reactant concentrations. Relaxation times over a wide range of concentrations were gathered, and the data were fit by a minimal mechanism in which a rapid bimolecular association step is followed by two sequential isomerization steps. The six rate constants for this mechanism were extracted from our data by relaxation analysis. The values determined for the six rate constants may be combined to calculate an overall equilibrium constant that is in excellent agreement with that obtained by independent equilibrium measurements. Additional stopped-flow experiments, using first sodium dodecyl sulfate to dissociate bound drug and second pseudo-first-order conditions to study the fast bimolecular step, provide independent verification of three of the six rate constants. The temperature dependence of four of the six rate constants was measured, allowing estimates of the activation energy of some of the steps to be made. We speculate that the three steps in the proposed mechanism may correspond to a rapid "outside" binding of daunomycin to DNA, followed by intercalation of the drug, followed by either conformational adjustment of the drug or DNA binding site or redistribution of bound drug to preferred sites.  相似文献   

17.
R A Alberty  R N Goldberg 《Biochemistry》1992,31(43):10610-10615
The criterion for chemical equilibrium at specified temperature, pressure, pH, concentration of free magnesium ion, and ionic strength is the transformed Gibbs energy, which can be calculated from the Gibbs energy. The apparent equilibrium constant (written in terms of the total concentrations of reactants like adenosine 5'-triphosphate, rather than in terms of species) yields the standard transformed Gibbs energy of reaction, and the effect of temperature on the apparent equilibrium constant at specified pressure, pH, concentration of free magnesium ion, and ionic strength yields the standard transformed enthalpy of reaction. From the apparent equilibrium constants and standard transformed enthalpies of reaction that have been measured in the adenosine 5'-triphosphate series and the dissociation constants of the weak acids and magnesium complexes involved, it is possible to calculate standard Gibbs energies of formation and standard enthalpies of formation of the species involved at zero ionic strength. This requires the convention that the standard Gibbs energy of formation and standard enthalpy of formation for adenosine in dilute aqueous solutions be set equal to zero. On the basis of this convention, standard transformed Gibbs energies of formation and standard transformed enthalpies of formation of adenosine 5'-trisphosphate, adenosine 5'-diphosphate, adenosine 5'-monophosphate, and adenosine at 298.15 K, 1 bar, pH = 7, a concentration of free magnesium ions of 10(-3) M, and an ionic strength of 0.25 M have been calculated.  相似文献   

18.
The equilibrium constant of the phosphoglyceromutase reaction was determined over a range of pH (5.4-7.9), in solutions of different ionic strength (0.06-0.3) and in the presence of Mg(2+), at 30 degrees C and at 20 degrees C. The values obtained (8.65-11.65) differ substantially from previously published values. The third acid dissociation constants were redetermined for 2- and 3-phosphoglycerate, and in contrast with previous reports the pK values (7.03 and 6.97 respectively at zero ionic strength) were closely similar. The Mg(2+)-binding constants were measured spectrophotometrically and the values, 286mm(-1) and 255mm(-1) for 2- and 3-phosphoglycerate at pH7 and ionic strength 0.02, were also very similar. From the relative lack of effect of temperature, pH and ionic strength it is concluded that the equilibrium constant differs from unity largely because of entropic factors. At low ionic strength, in the neutral region, the pH-dependence can be attributed to the small difference in the acid dissociation constants, but the difference in dissociation constants does not explain the pH-dependence in the acid region or at high ionic strength. Within physiological ranges of pH, Mg(2+) concentration and ionic strength there will be little variation in equilibrium constant.  相似文献   

19.
D-Glyceraldehyde 3-phosphate forms adducts with thiols. These adducts, which are presumed to be hemithioacetals, equilibrate rapidly with the unhydrated form of the aldehyde, which is the subtrate for D-glyceraldehyde 3-phosphate dehydrogenase. The adduct provides a substrate buffer system whereby a constant low free aldehyde concentration can be maintained during the oxidation of aldehyde by the enzyme and NAD+. With this system, the kinetics of the association of the aldehyde with the enzyme were examined. The rate profile for this reaction is a single exponential process, showing that all four active sites of the enzyme have equivalent and independent reactivity towards the aldehyde, with an apparent second-order rate constant of 5 X 10(7)M-1-S-1 at pH8.0 and 21 degrees C. The second-order rate constant becomes 8 X 10(7)M-1-S-1 when account is taken of the forward and reverse catalytic rate constants of the dehydrogenase. The pH-dependence of the observed rate constant is consistent with a requirement for the unprotonated form of a group of pK 6.1, which is the pK observed for second ionization of glyceraldehyde 3-phosphate. The rate of phosphorolysis of the acyl-enzyme intermediate during the steady-state oxidative phosphorylation of the aldehyde was studied, and is proportional to the total Pi concentration up to at least 1 mM-Pi at pH 7.5. The pH-dependence of the rate of NADH generation under these conditions can be explained by the rate law d[NADA]/dt = k[acy] holoenzyme][PO4(3-)-A1, where thioester bond, although kinetically indistinguishable rate equations for the reaction are possible. The rates of the phosphorolysis reaction and of the aldehyde-association reaction decrease with increasing ionic strength, suggesting that the active site of the enzyme has cationic groups which are involved in the reaction of the enzyme with anionic substrates.  相似文献   

20.
1. The pH and ionic strength dependence of the interaction of FMN with apoflavodoxin has been studied by fluorometry in the pH region 2-5, at 22 degrees C. 2. The rate constant of dissociation and the dissociation constant were experimentally determined; the rate constants of association were claculated at a given pH value. These constants depend on the ionic strength. The plots of these constants against the square root of the ionic strength are straight. 3. Our data have been interpreted in terms of the Br?nsted theory, which relates chemical reaction rates to ionic strength. The data indicate that the apoenzyme reaches its maximum net positive charge at pH 2.0-2.6. The calculated net charge in this pH region is between 11 and 12 and is in agreement with the theoretical value of 12 as deduced from the primary structure of the protein. The isoelectric point of the holoenzyme is about 4. 4. The rate constant of association extrapolated to zero ionic strength is 3.2-10(5)M-1-s-1 and is pH-independent. 5. The rate constant of dissociation and the dissociation constant extrapolated to zero ionic strength depend on the pH. The results are explained by assuming that there are two protein ionizations with a pK value of 3.4; these ionizing groups are possibly close to the FMN binding site.  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