首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
A C Storer  P R Carey 《Biochemistry》1985,24(24):6808-6818
The kinetic constants for the papain-catalyzed hydrolysis of the methyl thiono esters of N-benzoylglycine and N-(beta-phenylpropionyl)glycine are compared with those for the corresponding methyl ester substrates. The k2/Ks values for the thiono esters are 2-3 times higher than those for the esters, and both show bell-shaped pH dependencies with similar pKa's (approximately 4 and 9). The k3 values for the thiono esters are 30-60 times less than those for the esters and do not exhibit a pH dependency. Solvent deuterium isotope effects on k2/Ks and k3 were measured for the ester and thiono ester substrates of both glycine derivatives. Each thiono ester substrate showed an isotope effect similar to that for the corresponding ester substrate. Moreover, use of the proton inventory technique indicated that, as for esters, one proton is transferred in the transition state for deacylation during reactions involving thiono esters and the degree of heavy atom reorganization in the transition state is very similar in both cases. The k3 values for the hydrolysis of a series of para-substituted N-benzoylglycine esters were found to correlate with the k3 values for the corresponding para-substituted thiono esters [Carey, P. R., Lee, H., Ozaki, Y., & Storer, A. C. (1984) J. Am. Chem. Soc. 106, 8258-8262], showing that the rate-determining step for the deacylation of both thiolacyl and dithioacyl enzymes probably involves the disruption of a contact between the substrate's glycinic nitrogen atom and the sulfur of cysteine-25. It is concluded that the hydrolysis of esters and thiono esters proceeds by essentially the same reaction pathway. Due to an oxygen-sulfur exchange process the product released in the case of the N-(beta-phenylpropionyl)glycine thiono ester substrate is the dioxygen acid; however, for the N-benzoylglycine thiono ester substrate, the thiol acid is the initial product. This thiol acid then acts as a substrate for papain and reacylates the enzyme to eventually give the dioxygen acid product. It is shown that thiol acids are excellent substrates for papain.  相似文献   

2.
The lysyl ester analogue p-nitrophenyl alpha-amino-p-toluate hydrobromide was synthesized, and its reactions with thrombin, trypsin, and plasmin were studied by stopped-flow and conventional methods. Kinetic parameters were compared with those determined for the arginyl ester analogue, p-nitrophenyl p-guanidinobenzoate hydrochloride, with these enzymes. By following nitrophenol release or proflavin absorption changes in the stopped-flow spectrophotometer, the constants Ks (enzyme-substrate binding), k2 (acylation), and k3 (deacylation) were determined. The major findings were: (1) Ks values were similar regardless of the substrate or the enzyme; (2) k3 was approximately the same for the reaction of the lysyl ester analogue with any enzyme; (3) k2 for the lysyl ester analogue was 1100 times greater with trypsin than with thrombin; and (4) k2 with thrombin was 60 times greater for the arginyl than for the lysyl ester analogue. The results suggest that the limited cleavage of lysyl bonds by thrombin is due in part to restricted acylation rather than substrate binding. The active site of thrombin, compared with that of trypsin, appears to have a more stringent requirement for the spatial relationship between the cationic group and the bond cleaved in substrates.  相似文献   

3.
Two enzymes that catalyze the hydrolysis of l-asparagine have been isolated from extracts of Pseudomonas geniculata. After initial salt fractionation, the enzymes were separated by chromatography on diethylaminoethyl-Sephadex and purified to homogeneity by gel filtration, ion-exchange chromatography, and preparative polyacrylamide electrophoresis. The enzymes differ markedly in physicochemical properties. One enzyme, termed asparaginase A, has a molecular weight of approximately 96,000 whereas the other, termed asparaginase AG, has a molecular weight of approximately 135,000. Both enzymes are tetrameric. The asparaginase A shows activity only with l-asparagine as substrate, whereas the asparaginase AG hydrolyzes l-asparagine and l-glutamine at approximately equal rates and it is also active with d-asparagine and d-glutamine as substrates. The asparaginase A was found to be devoid of antitumor activity in mice, whereas the asparaginase AG was effective in increasing the mean survival times of both C3H mice carrying the asparagine-requiring Gardner 6C3HED tumor line and Swiss mice bearing the glutamine-requiring Ehrlich ascites tumor line. These differences in antitumor activity were related to differences in the K(m) values for l-asparagine for the two enzymes. The asparaginase A has a K(m) value of 1 x 10(-3) M for this substrate whereas the corresponding value for the AG enzyme is 1.5 x 10(-5) M. Thus the concentration of asparagine necessary for maximal activity of the asparaginase A is very high compared with that of the normal plasma level of asparagine, which is approximately 50 muM.  相似文献   

4.
In recent investigations we were able to demonstrate that the NADP-dependent aldehyde dehydrogenase of Acinetobacter calcoaceticus is an inducible enzyme localized in intracytoplasmic membranes limiting alkane inclusions. Long-chain aliphatic hydrocarbons and alkanols are inducers of the enzyme. It was purified by us and now kinetically characterized using the enzyme-micelle form, which contains bacterial phospholipids and a detergent (sodium cholate), too. The pH optimum of aldehyde dehydrogenase was determined to be at pH 10. The enzyme showed substrate inhibition (by aldehyde excess). The Ks and Km values of the leading substrate NADP+ were found to be 8.6 X 10(-5) and 10.3 X 10(-5)M independent of the chain-length of the aldehydes. The Km values of the aldehydes decreased depending on increasing chain-length (butanal: 1.6 X 10(-3), decanal: 1.5 X 10(-6)M). The Ki values (for inhibition by aldehyde excess) showed a similar behaviour (butanal: 7.5 X 10(-3), decanal: 3.5 X 10(-5)M) as well as the optimal aldehyde concentrations inducing the "maximal" reaction velocity (butanal: 5mM, decanal: 6 microM). The number of inhibiting aldehyde molecules per enzyme-substrate complex was determined to be n = 1. NADPH showed product inhibition kinetics (Ki(NADPH) = 2.2 X 10(-4)M), fatty acids did not. We were unable to measure a reverse reaction. The following ions and organic compounds were non-competitive inhibitors of the enzyme: Sn2+, Fe2+, Cu2+, BO3(3-), CN-, EDTA, o-phenanthroline, p-chloromercuri-benzoate, mercaptoethanol, phenylmethylsulfonyl fluoride, and diisopropylfluorophosphate; iodoacetate did not influence enzyme activity. Chloral hydrate was a competitive inhibitor of the aldehydes. Ethyl butyrate activates the enzyme, dependent on the chain-length of the aldehyde substrates.  相似文献   

5.
By solving simultaneously the equation for ''uniform binding'' [Albery & Knowles (1976) Biochemistry 15, 5631-5640] and the equation for ''differential binding'' [Chin (1983) J. Am. Chem. Soc. 105, 6502-6503], I derived the following simple equation for perfect enzymes (with single substrate and single product) under irreversible conditions: K2 = beta(1 + Rs)/1-beta(1 + Rs) where K2 is the internal equilibrium constant and beta is the Brönsted coefficient of the elementary catalytic step, and Rs is defined as [S]0/Ks, with [S]0 being the physiological substrate concentration and Ks being the substrate dissociation constant. The equation suggests that the perfect enzyme can have different internal thermodynamic properties depending on physiological conditions.  相似文献   

6.
Intact cells of Pseudomonas cepacia G4 completely degraded trichloroethylene (TCE) following growth with phenol. Degradation kinetics were determined for both phenol, used to induce requisite enzymes, and TCE, the target substrate. Apparent Ks and Vmax values for degradation of phenol by cells were 8.5 microM and 466 nmol/min per mg of protein, respectively. At phenol concentrations greater than 50 microM, phenol degradation was inhibited, yielding an apparent second-order inhibitory value, KSI, of 0.45 mM as modeled by the Haldane expression. A partition coefficient for TCE was determined to be 0.40 +/- 0.02, [TCEair]/[TCEwater], consistent with Henry's law. To eliminate experimental problems associated with TCE volatility and partitioning, a no-headspace bottle assay was developed, allowing for direct and accurate determinations of aqueous TCE concentration. By this assay procedure, apparent Ks and Vmax values determined for TCE degradation by intact cells were 3 microM and 8 nmol/min per mg of protein, respectively. Following a transient lag period, P. cepacia G4 degraded TCE at concentrations of at least 300 microM with no apparent retardation in rate. Consistent with Ks values determined for degradation, TCE significantly inhibited phenol degradation.  相似文献   

7.
Intact cells of Pseudomonas cepacia G4 completely degraded trichloroethylene (TCE) following growth with phenol. Degradation kinetics were determined for both phenol, used to induce requisite enzymes, and TCE, the target substrate. Apparent Ks and Vmax values for degradation of phenol by cells were 8.5 microM and 466 nmol/min per mg of protein, respectively. At phenol concentrations greater than 50 microM, phenol degradation was inhibited, yielding an apparent second-order inhibitory value, KSI, of 0.45 mM as modeled by the Haldane expression. A partition coefficient for TCE was determined to be 0.40 +/- 0.02, [TCEair]/[TCEwater], consistent with Henry's law. To eliminate experimental problems associated with TCE volatility and partitioning, a no-headspace bottle assay was developed, allowing for direct and accurate determinations of aqueous TCE concentration. By this assay procedure, apparent Ks and Vmax values determined for TCE degradation by intact cells were 3 microM and 8 nmol/min per mg of protein, respectively. Following a transient lag period, P. cepacia G4 degraded TCE at concentrations of at least 300 microM with no apparent retardation in rate. Consistent with Ks values determined for degradation, TCE significantly inhibited phenol degradation.  相似文献   

8.
In an effort to determine whether the Na+-dependent Pi transport system of Ehrlich ascites tumor cells exhibits specificity for H2PO4- or HPO4(-2), Pi fluxes were determined by measuring 32Pi-Pi self-exchange. Three experimental approaches were employed. First, the effect of pH on steady-state Pi transport at 0.5 and 5 mM was studied. Second, the relationship between Pi transport and Pi concentration (0.25-9.2 mM) at pH 5.6 and 7.9 was determined. Third, the dependence of Pi transport on [H2PO4-] (0.05-4.2 mM) at constant [HPO4(-2)] (0.5 mM), and the converse, [HPO4(-2)] (0.06-4.5 mM) at constant [H2PO4-] (0.5 mM), was evaluated. Ks (apparent half-saturation constant) and Jmax (maximal transport rate) were calculated by two methods: weighted linear regression (WLR) and a nonparametric procedure. The dependence of Pi flux on pH indicates that optimum transport occurs at pH 6.9. Pi transport decreases as pH is reduced when extracellular Pi is either 0.5 or 5 mM. However, at pH 7.9, Pi flux is reduced only in 0.5 mM Pi. At pH 5.6, H2PO4- comprises 93% of the total Pi present, and the calculated Ks is 0.055 +/- 0.026 mM (WLR). This is the same as the Ks determined from the initial phase of the flux vs. [H2PO4-] relationship (0.056 +/- 0.020 mM). However, at pH 7.9 (where 94% of Pi is HPO4(-2)), the measured Ks is 0.58 +/- 0.11 mM (WLR), which is ten times higher than at pH 5.6. This value is also five times greater than the Ks calculated from the flux vs. [HPO4(-20)] curve (0.106 +/- 0.16 mM). Kinetic parameters calculated by the nonparametric method, though somewhat different, gave similar relative results. Taken together, these results support two conclusions: (1) H2PO4- is the substrate for the Na+-dependent Pi transport system of the Ehrlich cell, and (2) H+ can inhibit Pi transport.  相似文献   

9.
The ATP-synthesis of the nuclear nucleoside-nucleotide phosphotransferase C was stimulated by progesterone alone and in combination with poly d(A-T) at 10(-10)-10(-9) M, by estradiol/poly d(A-T) at 5-10(-9) M, by cortisol alone and in combination with poly d(A-T) at 10(-9)-10(-8) M and by poly d(A-T) alone. No effect was observed using poly d(C-G). Using increasing adenosine/dTTP as substrates, cortisol, progesterone, estradiol/poly d(A-T) and poly d(A-T) alone caused positive cooperativity of the substrate saturation curves of the allosteric enzyme C by lowering the apparent Ks 0.5-values and by altering Vmax. The dTTP-synthesis of enzyme C was stimulated by both poly d(A-T) and poly d(C-G) alone and in combination with low estradiol (up to 10(-10) M)-and with higher progesterone concentrations (10(-8)-10(-7) M), whereas cortisol inhibited at higher concentrations completely. Using increasing thymidine/ATP as substrates progesterone and estradiol, also in combination with poly d(A-T) were positive effectors of the substrate saturation curves of enzyme C. By this, the apparent Ks 0.5-values or Vmax were changed. Cortisol could be shown to be a negative effector.  相似文献   

10.
Four enol lactones, bearing phenyl or 1-naphthyl substituents on the alpha or beta positions [3-phenyl-6-methylenetetrahydro-2-pyranone (alpha Ph6H, IIc), 3-(1-naphthyl)-6-methylenetetrahydro-2-pyranone (alpha Np6H, IId), 4-phenyl-6-methylenetetrahydro-2-pyranone (beta Ph6H, IIIc), and 4-(1-naphthyl)-6-methylenetetrahydro-2-pyranone (beta Np6H, IIId)], available as pure R and S enantiomers, have been studied as alternate substrate inhibitors of chymotrypsin. Kinetic constants for substrate binding (Ks) and acylation (ka) were determined by a competitive substrate assay, using succinyl-L-Ala-L-Ala-L-Pro-L-Phe p-nitroanilide; the deacylation rate constant (kd) was determined by the proflavin displacement assay. All lactones undergo rapid acylation (ka varies from 17 to 170 min-1) that shows little enantioselectivity; there is, however, pronounced enantioselectivity in substrate binding for three of the lactones [Ks(R/S) = 40-110]. In each case it is the enantiomer with the S configuration that has the higher affinity. In all cases, deacylation rates are slow, and in two cases, acyl enzymes with half-lives of 4.0 and 12.5 h at pH 7.2, 25 degrees C, are obtained (for beta Ph6H and alpha Np6H, respectively). In these cases, high deacylation enantioselectivity is observed [kd(S/R) = 60-70], and the lactone more weakly bound as a substrate (R enantiomer) gives the more stable acyl enzyme. Two hypotheses, involving hindrance of the attack of water or an exchange of the ester and ketone carbonyl groups in the acyl enzyme, are advanced as possible explanations for the high stability of these acyl enzymes.  相似文献   

11.
Bovine and dogfish chymotrypsins were compared to determine if chymotrypsin from a poikilothermic organism (spiny dogfish (Squalus acanthias] adapted to low temperatures possessed catalytic properties different from those of the same enzyme from a warm-blooded animal. An improved procedure was developed for isolating dogfish pancreatic chymotrypsin. The least hydrophobic and smallest substrate used, p-nitrophenyl acetate, had similar enthalpies of association (delta Ha) with both enzymes, whereas larger, more hydrophobic substrates had delta Ha values that were of opposite sign for the two enzymes. As the temperature increased, the association constants (1/Ks) for p-nitrophenyl valerate and p-nitrophenyltrimethyl acetate increased for dogfish chymotrypsin and decreased for bovine chymotrypsin, while the free energies of association (delta Ga) remained relatively constant. Acylation of chymotrypsin was 1.5-2.5 times slower in the dogfish enzyme than in the bovine enzyme except below 15 degrees C with p-nitrophenyltrimethyl acetate. delta H++ for acylation by p-nitrophenyltrimethyl acetate were 2.0 kcal/mol for the dogfish enzyme and 10.2 kcal/mol for the bovine, whereas delta H++ values were only slightly lower in the dogfish enzyme for the other two substrates. For all substrates, the deacylation rate constant (kcat) was greater with dogfish chymotrypsin than bovine. However, the free energies of activation (delta G++) for deacylation were nearly equal between the two enzymes for each of the substrates.  相似文献   

12.
The pH dependence of Vmax and Vmax/Km for hydrolysis of Dnp-Pro-Leu-Gly-Leu-Trp-Ala-D-Arg-NH2 at the Gly-Leu bond by porcine synovial collagenase and gelatinase was determined in the pH range 5-10. Both enzymes exhibited bell-shaped dependencies on pH for these two kinetic parameters, indicating that activity is dependent on at least two ionizable groups, one of which must be unprotonated and the other protonated. For collagenase, Vmax/Km data indicate that in the substrate-free enzyme, these groups have apparent pK values of 7.0 and 9.5, while the Vmax profile indicates similar pK values of 6.8 and 10.1 for the enzyme-substrate complex. The corresponding pH profiles of gelatinase were similar to those of collagenase, indicating the importance of groups with apparent pK values of 5.9 and 10.0 for the free enzyme and 5.9 and 11.1 for the enzyme-substrate complex. When these kinetic constants were determined in D2O using the peptide substrate, there was no significant effect on Vmax or Km for collagenase or Km for gelatinase. However, there was a deuterium isotope effect of approximately 1.5 on Vmax for gelatinase. These results indicate that a proton transfer step is not involved in the rate-limiting step for collagenase, but may be limiting with gelatinase. The Arrhenius activation energies for peptide bond hydrolysis of the synthetic peptide as well as the natural substrates were also determined for both enzymes. The activation energy (81 kcal) for hydrolysis of collagen by collagenase was nine times greater than that determined for the synthetic substrate (9.2 kcal). In contrast, the activation energy for hydrolysis of gelatin by gelatinase (26.3 kcal) was only 2.4 times greater than that for the synthetic substrate (11 kcal).  相似文献   

13.
1. Beta-Phenylpropionylthiocholine and N-(5-aminopentyl)-5-dimethylaminonaphthalene-1-sulphonamide (dansylcadaverine) serve as a pair of water-soluble (pH7.5) model substrates for transamidating enzymes. Amide formation could be followed directly through fluorescence measurements by monitoring the continuous extraction of the water-soluble coupling product, N-(beta-phenylpropionyl)dansylcadaverine, into n-heptane. By this procedure, the steady-state kinetics of glutamine-lysine endo-gamma-glutamyltransferase from human plasma (fibrinoligase, thrombin- and Ca2+-activated blood coagulation Factor XII) and from guinea-pig liver (liver transglutaminase) were investigated at 25 degrees C. 2. With beta-phenylpropionylthiocholine as the varied substrate, Lineweaver-Burk plots with various concentrations of dansylcadaverine intercept on the horizontal axis, suggesting that formation of the acyl-enzyme is rate limiting. 3. On the basis of functional normality of active sites, kcat. values of 1.8 s(-1) and 0.9 s(-1) were obtained for the plasma and liver gamma-glutamyltransferase respectively. The two enzymes show identical affinities for the first substrate, beta-phenylpropionylthiocholine, with Ka 4 times 10(-4) M. 4. Utilization of the second substrate, dansylcadaverine, appears to be an order of magnitude more efficient with the liver enzyme. 5. N-(5-Amino-3-thiapentyl)-5-dimethylaminonaphthalene-1-sulphonamide (dansylthiacadaverine) could be used instead of dansylcadaverine in the fluorescent kinetic system. 6. Competitive inhibition by a non-fluorescent amine substrate histamine was also evaluated.  相似文献   

14.
A theoretical investigation is presented which allows the calculation of states of maximal reaction rates for single enzymes and for unbranched enzymatic chains. As an extension to previous papers (Heinrich & Holzhütter, 1985, Biomed. biochim. Acta 44, 959-969; Heinrich et al., 1987, Bull. math. Biol. 49, 539-595) a detailed enzymatic mechanism was taken into consideration. Conclusions are drawn for the optimal values of the microscopic rate constants as well as of the maximal activities and Michaelis constants. Ten solutions are found which depend on the equilibrium constant as well as on the concentrations of substrates and products. It is shown that for high equilibrium constants one of the solutions applies to a very large range of the concentrations of the outer reactants. This solution is characterized by maximal values of the rate constants of all forward reactions and by non-maximal values of the rate constants of all backward reactions. In contrast to previous assumptions (Albery & Knowles, 1976b, Biochemistry 15, 5631-5640; Burbaum et al., 1989, Biochemistry 28, 9293-9305) states of maximal reaction rate are not always characterized by the highest possible values of the second-order rate constants which are related to the diffusion of the substrate and the product to the active site of the enzyme. Predictions are made concerning the ratios of maximal activities in optimal states as well as for the adaptation of the Michaelis constants to the concentrations of the outer reactants. Using metabolic control analysis it is shown that the solutions obtained for single enzymes may also be applied in multi-enzyme systems.  相似文献   

15.
F1,6BPases from porcine and bovine lung were isolated and their kinetic properties were determined. Ks, Kis and beta were determined assuming partial-noncompetitive inhibition (simple intersecting hyperbolic noncompetitive inhibition) of the enzyme by the substrate. Values for Ks were 4.1 and 4.4 microM for porcine and bovine F1,6BPase, respectively and values for 1 were close to 0.55 in both cases. Kis were 9 and 15 microM for porcine and bovine F1,6BPase, respectively. I0.5 for AMP were determined as 7 microM for pig enzyme and 14 microM for F1,6BPase from bovine lung. The enzymes were inhibited by F2,6BP with Ki's of 0.19 and 0.21 microM for porcine and bovine enzymes, respectively. In the presence of AMP concentration equal to I0.5, the Ki values for pig and bovine enzymes were 0.07 and 0.09 microM, respectively. The levels of F2,6BP, AMP and antioxidant enzymes activities in pig and bovine lung tissues were also determined. The cDNA coding sequence of pig lung F1,6BPase1 showed a high homology with pig liver enzyme, differing only in four positions (G/C-63, T/A-808, G/C-884 and T/A-1005) resulting in a single amino acid substitution (Gly-295 for Ala-295). It is hypothesized that the lung F1,6BPase participates in gluconeogenesis, surfactant synthesis and antioxidant reactions.  相似文献   

16.
Albumin is generally regarded as an inert protein with no enzyme activity. However, albumin has esterase activity as well as aryl acylamidase activity. A new acetanilide substrate, o-nitrotrifluoroacetanilide (o-NTFNAC), which is more reactive than the classical o-nitroacetanilide, made it possible to determine the catalytic parameters for hydrolysis by fatty-acid free human serum albumin. Owing to the low enzymatic activity of albumin, kinetic studies were performed at high albumin concentration (0.075 mM). The albumin behavior with this substrate was Michaelis-Menten like. Kinetic analysis was performed according to the formalism used for catalysis at high enzyme concentration. This approach provided values for the turnover and dissociation constant of the albumin-substrate complex: k(cat) = 0.13 +/- 0.02 min(-1) and Ks = 0.67 +/- 0.04 mM. MALDI-TOF experiments showed that unlike the ester substrate p-nitrophenyl acetate, o-NTFNAC does not form a stable adduct (acetylated enzyme). Kinetic analysis and MALDI-TOF experiments demonstrated that hydrolysis of o-NTFNAC by albumin is fully rate-limited by the acylation step (k(cat) = k2). Though the aryl acylamidase activity of albumin is low (k(cat)/Ks = 195 M(-1)min(-1)), because of its high concentration in human plasma (0.6-1 mM), albumin may participate in hydrolysis of aryl acylamides through second-order kinetics. This suggests that albumin may have a role in the metabolism of endogenous and exogenous aromatic amides, including drugs and xenobiotics.  相似文献   

17.
Bestatin, [(2S,3R)-3-amino-2-hydroxy-4-phenyl-butanoyl]-L-leucine, a known inhibitor of aminopeptidases, is shown to be a potent linear competitive inhibitor (KI,2.7 nM) of a dipeptidase purified from Ehrlich-Lettré hyperdiploid mouse ascites tumor cells. This inhibition can be classified as "slow binding" but not as "tight binding." Substrate protects the enzyme from bestatin inhibition when enzyme and inhibitor are in approximately equimolar concentrations. Addition of substrate (6 mM) partially (by about 20%) reverses dipeptidase inhibition by bestatin, but the time required for maximum recovery depends on the nature of the substrate. Substrates with lower Km (0.28-1.4 mM) values that exhibit substantial substrate inhibition require longer times (23-65 min) than those with higher Km values that show little substrate inhibition. Substrates with Km values higher than 1.5 mM do not reverse inhibition. The inhibition of the tumor dipeptidase by bestatin has been compared with inhibition by a variety of inhibitors of other Zn-metallo-proteolytic enzymes. These inhibitors were far less potent (KI, 0.063-10 mM), indicating a difference between the tumor dipeptidase and other enzymes of that class. Our results are discussed in terms of a postulated model of the bestatin molecule in the active site of the tumor dipeptidase, an enzyme which has not been studied by x-ray crystallographic means. The phenyl group of bestatin is placed in a hydrophobic pocket that is external but adjacent to the active site of the tumor dipeptidase. The shape of this pocket, as it appears from our results plus modeling, is such that only certain R groups of substrate can fit. The existence of such a pocket might explain the differential effect of substrates in the reversal of bestatin inhibition of the dipeptidase and also might explain substrate inhibition by misalignment of R groups into this pocket.  相似文献   

18.
Pre-steady-state and steady-state kinetics of the papain (EC 3.4.22.2)-catalyzed hydrolysis of N-alpha-carbobenzoxyglycine p-nitrophenyl ester (ZGlyONp) have been determined between pH 3.0 and 9.5 (I = 0.1 M) at 21 +/- 0.5 degrees C. The results are consistent with the minimum three-step mechanism involving the acyl X enzyme intermediate E X P: (Formula: see text). The formation of the E X S complex may be regarded as a rapid pseudoequilibrium process; the minimum values for k+1 are 8.0 microM-1 s-1 (pH less than or equal to 3.5) and 0.40 microM-1 s-1 (pH greater than 6.0), and that for k-1 is 600 s-1 (pH independent). The pH profile of k+2/Ks (= kcat/Km; Ks = k-1/k+1) reflects the ionization of two groups with pK' values of 4.5 +/- 0.1 and 8.80 +/- 0.15 in the free enzyme. The pH dependence of k+2 and k+3 (measured only at pH values below neutrality) implicates one ionizing group in the acylation and deacylation step with pK' values of 5.80 +/- 0.15 and 3.10 +/- 0.15, respectively. As expected from the pH dependences of k+2/Ks (= kcat/Km) and k+2, the value of Ks changes with pH from 7.5 X 10(1) microM (pH less than or equal to 3.5) to 1.5 X 10(3) microM (pH greater than 6.0). Values of k-2 and k-3 are close to zero over the whole pH range explored (3.0 to 9.5). The pH dependence of kinetic parameters indicates that at acid pH values (less than or equal to 3.5), the k+2 step is rate limiting in catalysis, whereas for pH values higher than 3.5, k+3 becomes rate limiting. The observed ionizations probably reflect the acid-base equilibria of residues involved in the catalytic diad of papain, His159-Cys25. Comparison with catalytic properties of ficins and bromelains suggests that the results reported here may be of general significance for cysteine proteinase catalyzed reactions.  相似文献   

19.
M Philipp  I H Tsai  M L Bender 《Biochemistry》1979,18(17):3769-3773
The p-nitrophenyl esters of straight-chain fatty acids were used as substrates of the enzyme subtilisin Novo (EC 3.4.4.16) and its chemically produced artificial enzyme thiolsubtilisin. Subtilisin and thiolsubtilisin pH--activity profiles were determined, and kinetic effects of the active site O-S substitution were observed. Among the substrates tested, both enzymes show highest specificity with p-nitrophenyl butyrate. It was also found that subtilisin is more sensitive to changes in substrate chain length than is thiolsubtilisin. Second-order acylation rate constants (k2/Ks) are remarkably similar for both enzymes. However, thiolsubtilisin deacylation rate constants and Km values are lower than analogous subtilisin constants. While thiolsubtilisin deacylation rate constants give a pH profile identical with that of subtilisin, the pH profile of thiolsubtilisin acylation rate constants shows an active site pK value lowered from the subtilisin pK of 7.15 and exhibits an inflection point at pH 8.45, which is absent in subtilisin.  相似文献   

20.
Dihydroorotase (DHOase, EC 3.5.2.3) from the extreme thermophile Bacillus caldolyticus has been subcloned, sequenced, expressed, and purified as a monomer. The catalytic properties of this thermophilic DHOase have been compared with another type I enzyme, the DHOase domain from hamster, to investigate how the thermophilic enzyme is adapted to higher temperatures. B. caldolyticus DHOase has higher Vmax and Ks values than hamster DHOase at the same temperature. The thermodynamic parameters for the binding of L-dihydroorotate were determined at 25 degrees C for hamster DHOase (deltaG = -6.9 kcal/mol, deltaH = -11.5 kcal/mol, TdeltaS = -4.6 kcal/mol) and B. caldolyticus DHOase (deltaG = -5.6 kcal/mol, deltaH = -4.2 kcal/mol, TdeltaS = +1.4 kcal/mol). The smaller enthalpy release and positive entropy for thermophilic DHOase are indicative of a weakly interacting Michaelis complex. Hamster DHOase has an enthalpy of activation of 12.3 kcal/mol, similar to the release of enthalpy upon substrate binding, rendering the kcat/Ks value almost temperature independent. B. caldolyticus DHOase shows a decrease in the enthalpy of activation from 12.2 kcal/mol at temperatures from 30 to 50 degrees C to 5.3 kcal/mol for temperatures of 50-70 degrees C. Vibrational energy at higher temperatures may facilitate the transition ES --> ES(double dagger), making kcat/Ks almost temperature independent. The pseudo-first-order rate constant for water attack on L-dihydroorotate, based on experiments at elevated temperature, is 3.2 x 10(-11) s(-1) at 25 degrees C, with deltaH(double dagger) = 24.7 kcal/mol and TdeltaS(double dagger) = -6.9 kcal/mol. Thus, hamster DHOase enhances the rate of substrate hydrolysis by a factor of 1.6 x 10(14), achieving this rate enhancement almost entirely by lowering the enthalpy of activation (delta deltaH(double dagger) = -19.5 kcal/mol). Both the rate enhancement and transition state affinity of hamster DHOase increase steeply with decreasing temperature, consistent with the development of H-bonds and electrostatic interactions in the transition state that were not present in the enzyme-substrate complex in the ground state.  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