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 62 毫秒
1.
Catalytic reaction pathway for the mitogen-activated protein kinase ERK2   总被引:2,自引:0,他引:2  
Prowse CN  Hagopian JC  Cobb MH  Ahn NG  Lew J 《Biochemistry》2000,39(20):6258-6266
The structural, functional, and regulatory properties of the mitogen-activated protein kinases (MAP kinases) have long attracted considerable attention owing to the critical role that these enzymes play in signal transduction. While several MAP kinase X-ray crystal structures currently exist, there is by comparison little mechanistic information available to correlate the structural data with the known biochemical properties of these molecules. We have employed steady-state kinetic and solvent viscosometric techniques to characterize the catalytic reaction pathway of the MAP kinase ERK2 with respect to the phosphorylation of a protein substrate, myelin basic protein (MBP), and a synthetic peptide substrate, ERKtide. A minor viscosity effect on k(cat) with respect to the phosphorylation of MBP was observed (k(cat) = 10 +/- 2 s(-1), k(cat)(eta) = 0.18 +/- 0.05), indicating that substrate processing occurs via slow phosphoryl group transfer (12 +/- 4 s(-1)) followed by the faster release of products (56 +/- 4 s(-1)). At an MBP concentration extrapolated to infinity, no significant viscosity effect on k(cat)/K(m(ATP)) was observed (k(cat)/K(m(ATP)) = 0.2 +/- 0.1 microM(-1) s(-1), k(cat)/K(m(ATP))(eta) = -0.08 +/- 0.04), consistent with rapid-equilibrium binding of the nucleotide. In contrast, at saturating ATP, a full viscosity effect on k(cat)/K(m) for MBP was apparent (k(cat)/K(m(MBP)) = 2.4 +/- 1 microM(-1) s(-1), k(cat)/K(m(MBP))(eta) = 1.0 +/- 0.1), while no viscosity effect was observed on k(cat)/K(m) for the phosphorylation of ERKtide (k(cat)/K(m(ERKtide)) = (4 +/- 2) x 10(-3) microM(-1) s(-1), k(cat)/K(m(ERKtide))(eta) = -0.02 +/- 0.02). This is consistent with the diffusion-limited binding of MBP, in contrast to the rapid-equilibrium binding of ERKtide, to form the ternary Michaelis complex. Calculated values for binding constants show that the estimated value for K(d(MBP)) (/= 1.5 mM). The dramatically higher catalytic efficiency of MBP in comparison to that of ERKtide ( approximately 600-fold difference) is largely attributable to the slow dissociation rate of MBP (/=56 s(-1)), from the ERK2 active site.  相似文献   

2.
Stewart RC  Jahreis K  Parkinson JS 《Biochemistry》2000,39(43):13157-13165
The histidine protein kinase CheA plays a central role in the bacterial chemotaxis signal transduction pathway. Autophosphorylated CheA passes its phosphoryl group to CheY very rapidly (k(cat) approximately 750 s(-)(1)). Phospho-CheY in turn influences the direction of flagellar rotation. The autophosphorylation site of CheA (His(48)) resides in its N-terminal P1 domain. The adjacent P2 domain provides a high-affinity binding site for CheY, which might facilitate the phosphotransfer reaction by tethering CheY in close proximity to the phosphodonor located in P1. To explore the contribution of P2 to the CheA --> CheY phosphotransfer reaction in the Escherichia coli chemotaxis system, we examined the transfer kinetics of a mutant CheA protein (CheADeltaP2) in which the 98 amino acid P2 domain had been replaced with an 11 amino acid linker. We used rapid-quench and stopped-flow fluorescence experiments to monitor phosphotransfer to CheY from phosphorylated wild-type CheA and from phosphorylated CheADeltaP2. The CheADeltaP2 reaction rates were significantly slower and the K(m) value was markedly higher than the corresponding values for wild-type CheA. These results indicate that binding of CheY to the P2 domain of CheA indeed contributes to the rapid kinetics of phosphotransfer. Although phosphotransfer was slower with CheADeltaP2 (k(cat)/K(m) approximately 1.5 x 10(6) M(-)(1) s(-)(1)) than with wild-type CheA (k(cat)/K(m) approximately 10(8) M(-)(1) s(-)(1)), it was still orders of magnitude faster than the kinetics of CheY phosphorylation by phosphoimidazole and other small molecule phosphodonors (k(cat)/K(m) approximately 5-50 M(-)(1) s(-)(1)). We conclude that the P1 domain of CheA also makes significant contributions to phosphotransfer rates in chemotactic signaling.  相似文献   

3.
Rho-associated coiled-coil kinase, ROCK, is implicated in Rho-mediated cell adhesion and smooth muscle contraction. Animal models suggest that the inhibition of ROCK can ameliorate conditions, such as vasospasm, hypertension, and inflammation. As part of our effort to design novel inhibitors of ROCK, we investigated the kinetic mechanism of ROCK I. Steady-state bisubstrate kinetics, inhibition kinetics, isotope partition analysis, viscosity effects, and presteady-state kinetics were used to explore the kinetic mechanism. Plots of reciprocals of initial rates obtained in the presence of nonhydrolyzable ATP analogues and the small molecule inhibitor of ROCK, Y-27632, against the reciprocals of the peptide concentrations yielded parallel lines (uncompetitive pattern). This pattern is indicative of an ordered binding mechanism, with the peptide adding first. The staurosporine analogue K252a, however, gave a noncompetitive pattern. When a pulse of (33)P-gamma-ATP mixed with ROCK was chased with excess unlabeled ATP and peptide, 0.66 enzyme equivalent of (33)P-phosphate was incorporated into the product in the first turnover. The presence of ATPase activity coupled with the isotope partition data is a clear evidence for the existence of a viable [E-ATP] complex in the kinase reaction and implicates a random binding mechanism. The k(cat)/K(m) parameters were fully sensitive to viscosity (viscosity effects of 1.4 +/- 0.2 and 0.9 +/- 0.3 for ATP and peptide 5, respectively), and therefore, the barriers to dissociation of either substrate are higher than the barrier for the phosphoryl transfer step. As a consequence, not all the binding steps are at fast equilibrium. The observation of a burst in presteady-state kinetics (k(b) = 10.2 +/- 2.1 s(-)(1)) and the viscosity effect on k(cat) of 1.3 +/- 0.2 characterize the phosphoryl transfer step to be fast and the release of product and/or the enzyme isomerization step accompanying it as rate-limiting at V(max) conditions. From the multiple kinetic studies, most of the rate constants for the individual steps were either evaluated or estimated.  相似文献   

4.
In the present study, the enzymatic mechanism of ERK2 is re-examined by a combination of steady-state kinetic studies in the absence and presence of viscosogenic agents. Kinetic studies carried out in various concentrations of sucrose revealed that both k(cat) and k(cat)/K(m) for either ATP or EtsDelta138 were highly sensitive to solvent viscosity, suggesting that the rapid equilibrium assumption is not valid for the phosphorylation of protein substrate by ERK2. Furthermore, the kinetic analysis with the minimal random Bi Bi reaction mechanism is shown to be inconsistent with the principle of the detailed balance. This inconsistent calculation strongly suggests that there is isomerization of the enzyme-substrate ternary complex. The viscosity-dependent steady-state kinetic data are combined to establish a kinetic mechanism for the ERK2-catalyzed reaction that predicts initial reaction velocities under varying concentrations of ATP and substrate. These results complement previous structure-function studies of mitogen-activated protein kinases and provide important insight for mechanistic interpretation of the kinase functions.  相似文献   

5.
Yao L  Li Y  Wu Y  Liu A  Yan H 《Biochemistry》2005,44(15):5940-5947
Yeast cytosine deaminase (yCD), a zinc metalloenzyme, catalyzes the hydrolytic deamination of cytosine to uracil. The enzyme is of great biomedical interest because it also catalyzes the deamination of the prodrug 5-fluorocytosine (5FC) to form the anticancer drug 5-fluorouracil (5FU). yCD/5FC is one of the most widely used enzyme/prodrug combinations for gene-directed enzyme prodrug therapy for the treatment of cancers. A pH indicator assay has been developed for the measurement of the steady-state kinetic parameters for the deamination reaction. Transient kinetic studies have shown that the product release is a rate-limiting step in the activation of the prodrug 5FC by yCD. The rate constant of the chemical step for the forward reaction (250 s(-)(1)) is approximately 8 times that of the product release (31 s(-)(1)) and approximately 15 times k(cat) (17 s(-)(1)). The transient kinetic results are consistent with those of the steady-state kinetic analysis in the sense that the k(cat) and K(m) values calculated from the rate constants determined by the transient kinetic analysis are in close agreement with those measured by the steady-state kinetic analysis. NMR experiments have demonstrated that free 5FU is in slow exchange with its complex with yCD but has a low affinity for yCD. The transient kinetic and NMR results together suggest that the release of 5FU is rate-limiting in the activation of the prodrug 5FC by yCD and may involve multiple steps.  相似文献   

6.
Waas WF  Rainey MA  Szafranska AE  Dalby KN 《Biochemistry》2003,42(42):12273-12286
Extracellular regulated protein kinase 2 (ERK2) is a eukaryotic protein kinase whose activity is regulated by mitogenic stimuli. To gain insight into the catalytic properties of ERK2 and to complement structure-function studies, we undertook a pre-steady state kinetic analysis of the enzyme. To do this, ERK2 was quantitatively activated by MAPKK1 in vitro by monitoring the stoichiometry and site specificity of phosphorylation using a combination of protein mass spectrometry, tryptic peptide analysis, and (32)P radiolabeling. Using a quench-flow apparatus, MgATP(2-) was rapidly mixed (<1 ms) with both ERK2 and the protein substrate EtsDelta138 in the presence of a saturating total concentration (20 mM) of magnesium ion at 27 degrees C and pH 7.5. An exponential burst of product was observed over the first few milliseconds that followed mixing. This burst had an amplitude alpha of 0.44 and was followed by a slower linear phase. The pre-steady state burst is consistent with two partially rate-limiting enzymatic steps, which have the following rate constants: k(2) = 109 +/- 9 s(-1) and k(3) = 56 +/- 4 s(-1). These are attributed to rapid phosphorylation of EtsDelta138 and the process of product release, respectively. Single-turnover experiments provided an independent determination of k(2) (106 +/- 25 s(-1)). The observed catalytic constant (k(cat)(obs)) was found to be sensitive to the concentration of ERK2. The data fit a model in which ERK2 monomers form dimers and suggest that both the monomeric and dimeric forms of ERK2 are active with catalytic constants (k(cat)) of 25 and 37 s(-1), respectively. In addition, the model suggests that in the presence of saturating concentrations of both magnesium and substrates ERK2 subunits dissociate with a dissociation constant (K(d)) of 32 +/- 16 nM.  相似文献   

7.
Klimacek M  Nidetzky B 《Biochemistry》2002,41(31):10158-10165
Mannitol dehydrogenases (MDH) are a family of Zn(2+)-independent long-chain alcohol dehydrogenases that catalyze the regiospecific NAD(+)-dependent oxidation of a secondary alcohol group in polyol substrates. pH and primary deuterium kinetic isotope effects on kinetic parameters for reaction of recombinant MDH from Pseudomonas fluorescens with D-mannitol have been measured in H(2)O and D(2)O at 25 degrees C and used to determine the relative timing of C-H and O-H bond cleavage steps during alcohol conversion. The enzymatic rates decreased at low pH; apparent pK values for log(k(cat)/K(mannitol)) and log k(cat) were 9.2 and 7.7 in H(2)O, respectively, and both were shifted by +0.4 pH units in D(2)O. Proton inventory plots for k(cat) and k(cat)/K(mannitol) were determined at pL 10.0 using protio or deuterio alcohol and were linear at the 95% confidence level. They revealed the independence of primary deuterium isotope effects on the atom fraction of deuterium in a mixed H(2)O-D(2)O solvent and yielded single-site transition-state fractionation factors of 0.43 +/- 0.05 and 0.47 +/- 0.01 for k(cat)/K(mannitol) and k(cat), respectively. (D)(k(cat)/K(mannitol)) was constant (1.80 +/- 0.20) in the pH range 6.0-9.5 and decreased at high pH to a limiting value of approximately 1. Measurement of (D)(k(cat)/K(fructose)) at pH 10.0 and 10.5 using NADH deuterium-labeled in the 4-pro-S position gave a value of 0.83, the equilibrium isotope effect on carbonyl group reduction. A mechanism of D-mannitol oxidation by MDH is supported by the data in which the partly rate-limiting transition state of hydride transfer is stabilized by a single solvation catalytic proton bridge. The chemical reaction involves a pH-dependent internal equilibrium which takes place prior to C-H bond cleavage and in which proton transfer from the reactive OH to the enzyme catalytic base may occur. Loss of a proton from the enzyme at high pH irreversibly locks the ternary complex with either alcohol or alkoxide bound in a conformation committed of undergoing NAD(+) reduction at a rate about 2.3-fold slower than the corresponding reaction rate of the protonated complex. Transient kinetic studies for D-mannitol oxidation at pH(D) 10.0 showed that the solvent isotope effect on steady-state turnover originates from a net rate constant of NADH release that is approximately 85% rate-limiting for k(cat) and 2-fold smaller in D(2)O than in H(2)O.  相似文献   

8.
M Slatner  B Nidetzky  K D Kulbe 《Biochemistry》1999,38(32):10489-10498
To characterize catalysis by NAD-dependent long-chain mannitol 2-dehydrogenases (MDHs), the recombinant wild-type MDH from Pseudomonas fluorescens was overexpressed in Escherichia coli and purified. The enzyme is a functional monomer of 54 kDa, which does not contain Zn(2+) and has B-type stereospecificity with respect to hydride transfer from NADH. Analysis of initial velocity patterns together with product and substrate inhibition patterns and comparison of primary deuterium isotope effects on the apparent kinetic parameters, (D)k(cat), (D)(k(cat)/K(NADH)), and (D)(k(cat)/K(fructose)), show that MDH has an ordered kinetic mechanism at pH 8.2 in which NADH adds before D-fructose, and D-mannitol and NAD are released in that order. Isomerization of E-NAD to a form which interacts with D-mannitol nonproductively or dissociation of NAD from the binary complex after isomerization is the slowest step (>/=110 s(-)(1)) in D-fructose reduction at pH 8.2. Release of NADH from E-NADH (32 s(-)(1)) is the major rate-limiting step in mannitol oxidation at this pH. At the pH optimum for D-fructose reduction (pH 7.0), the rate of hydride transfer contributes significantly to rate limitation of the catalytic cascade and the overall reaction. (D)(k(cat)/K(fructose)) decreases from 2.57 at pH 7.0 to a value of 相似文献   

9.
Jin Y  Penning TM 《Biochemistry》2006,45(43):13054-13063
Human type 3 3alpha-hydroxysteroid dehydrogenase, or aldo-keto reductase (AKR) 1C2, eliminates the androgen signal in human prostate by reducing 5alpha-dihydrotestosterone (DHT, potent androgen) to form 3alpha-androstanediol (inactive androgen), thereby depriving the androgen receptor of its ligand. The k(cat) for the NADPH-dependent reduction of DHT catalyzed by AKR1C2 is 0.033 s(-1). We employed transient kinetics and kinetic isotope effects to dissect the contribution of discrete steps to this low k(cat) value. Stopped-flow experiments to measure the formation of the AKR1C2.NADP(H) binary complex indicated that two slow isomerization events occur to yield a tight complex. A small primary deuterium isotope effect on k(cat) (1.5) and a slightly larger effect on k(cat)/K(m) (2.1) were observed in the steady state. In the transient state, the maximum rate constant for the single turnover of DHT (k(trans)) was determined to be 0.11 s(-1) for the NADPH-dependent reaction, which was approximately 4-fold greater than the corresponding k(cat) x k(trans) was significantly reduced when NADPD was substituted for NADPH, resulting in an apparent (D)k(trans) of 3.5. Thus, the effects of isotopic substitution on the hydride transfer step were masked by slow events that follow or precede the chemical transformation. Transient multiple-turnover reactions generated curvilinear reaction traces, consistent with the product formation and release occurring at comparable rates. Global fitting analysis of the transient kinetic data enabled the estimate of the rate constants for the three-step cofactor binding/release model and for the minimal ordered bi-bi turnover mechanism. Results were consistent with a kinetic mechanism in which a series of slow events, including the chemical step (0.12 s(-1)), the release of the steroid product (0.081 s(-1)), and the release of the cofactor product (0.21 s(-1)), combine to yield the overall observed low turnover number.  相似文献   

10.
Microsomal glutathione transferase 1 (MGST1) displays the unique ability to be activated, up to 30-fold, by the reaction with sulfhydryl reagents, e.g., N-ethylmaleimide. Analysis of glutathione (GSH) thiolate formation, which occurs upon mixing activated MGST1 with GSH, reveals biphasic kinetics, where the rapid phase dominated at higher GSH concentrations. The kinetic behavior suggests a two-step mechanism consisting of a rapid GSH-binding step (K(D)(GSH) approximately 10 mM), followed by slower formation of thiolate (k(2) approximately 10 s(-1)). The release rate (or protonation of the enzyme GSH thiolate complex) of GS(-) was slow (k(-2) = 0.016 s(-1)), consistent with overall tight binding of GSH. Electrophilic second substrates react rapidly with the E*GS(-) complex, and again, a two-step mechanism is suggested. In comparison to the unactivated enzyme [Morgenstern et al. (2001) Biochemistry 40, 3378-3384], the mechanisms of GSH thiolate formation and electrophile interaction are similar; however, thiolate anion formation is enhanced 30-fold in the activated enzyme, contributing to an increased k(cat) (3.6 s(-1)). Interestingly, in the activated enzyme, thiolate formation and proton release from the enzyme are not strictly coupled, because proton release (as well as k(cat)) was found to be approximately 4 times slower than GSH thiolate formation in an unbuffered system. Solvent kinetic isotope effect measurements demonstrated a 2-fold decrease in the rate constant (k(2)) for thiolate formation and k(cat) (in the reaction with 1-chloro-2,4-dinitrobenzene) for both unactivated and activated MGST1. This indicates that thiolate formation contributes to k(cat) for the activated enzyme, as suggested previously for unactivated MGST1. The stoichiometry of thiolate formation, proton release, and burst kinetics suggested utilization of one GSH molecule per enzyme trimer.  相似文献   

11.
The Staphylococcus aureus transpeptidase SrtA catalyzes the covalent attachment of LPXTG-containing virulence and colonization-associated proteins to cell-wall peptidoglycan in Gram-positive bacteria. Recent structural characterizations of staphylococcal SrtA, and related transpeptidases SrtB from S. aureus and Bacillus anthracis, provide many details regarding the active site environment, yet raise questions with regard to the nature of catalysis and active site cysteine thiol activation. Here we re-evaluate the kinetic mechanism of SrtA and shed light on aspects of its catalytic mechanism. Using steady-state, pre-steady-state, bisubstrate kinetic studies, and high-resolution electrospray mass spectrometry, revised steady-state kinetic parameters and a ping-pong hydrolytic shunt kinetic mechanism were determined for recombinant SrtA. The pH dependencies of kinetic parameters k(cat)/K(m) and k(cat) for the substrate Abz-LPETG-Dap(Dnp)-NH(2) were bell-shaped with pK(a) values of 6.3 +/- 0.2 and 9.4 +/- 0.2 for k(cat) and 6.2 +/- 0.2 and 9.4 +/- 0.2 for k(cat)/K(m). Solvent isotope effect (SIE) measurements revealed inverse behavior, with a (D)2(O)k(cat) of 0.89 +/- 0.01 and a (D)2(O)(k(cat)/K(m)) of 0.57 +/- 0.03 reflecting an equilibrium SIE. In addition, SIE measurements strongly implicated Cys184 participation in the isotope-sensitive rate-determining chemical step when considered in conjunction with an inverse linear proton inventory for k(cat). Last, the pH dependence of SrtA inactivation by iodoacetamide revealed a single ionization for inactivation. These studies collectively provide compelling evidence for a reverse protonation mechanism where a small fraction (ca. 0.06%) of SrtA is competent for catalysis at physiological pH, yet is highly active with an estimated k(cat)/K(m) of >10(5) M(-)(1) s(-)(1).  相似文献   

12.
The P450 2E1-catalyzed oxidation of ethanol to acetaldehyde is characterized by a kinetic deuterium isotope effect that increases K(m) with no effect on k(cat), and rate-limiting product release has been proposed to account for the lack of an isotope effect on k(cat) (Bell, L. C., and Guengerich, F. P. (1997) J. Biol. Chem. 272, 29643-29651). Acetaldehyde is also a substrate for P450 2E1 oxidation to acetic acid, and k(cat)/K(m) for this reaction is at least 1 order of magnitude greater than that for ethanol oxidation to acetaldehyde. Acetic acid accounts for 90% of the products generated from ethanol in a 10-min reaction, and the contribution of this second oxidation has been overlooked in many previous studies. The noncompetitive intermolecular kinetic hydrogen isotope effects on acetaldehyde oxidation to acetic acid ((H)(k(cat)/K(m))/(D)(k(cat)/K(m)) = 4.5, and (D)k(cat) = 1.5) are comparable with the isotope effects typically observed for ethanol oxidation to acetaldehyde, and k(cat) is similar for both reactions, suggesting a possible common catalytic mechanism. Rapid quench kinetic experiments indicate that acetic acid is formed rapidly from added acetaldehyde (approximately 450 min(-1)) with burst kinetics. Pulse-chase experiments reveal that, at a subsaturating concentration of ethanol, approximately 90% of the acetaldehyde intermediate is directly converted to acetic acid without dissociation from the enzyme active site. Competition experiments suggest that P450 2E1 binds acetic acid and acetaldehyde with relatively high K(d) values, which preclude simple tight binding as an explanation for rate-limiting product release. The existence of a rate-determining step between product formation and release is postulated. Also proposed is a conformational change in P450 2E1 occurring during the course of oxidation and the discrimination of P450 2E1 between acetaldehyde and its hydrated form, the gem-diol. This multistep P450 reaction is characterized by kinetic control of individual reaction steps and by loose binding of all ligands.  相似文献   

13.
Heat shock induces a stress response in mammalian cells and can also lead to apoptotic cell death. Here we report that a 36-kDa myelin basic protein (MBP) kinase detected by an in-gel kinase assay can be drastically activated in several cell types by heat shock. Immunoblot analysis revealed that this 36-kDa MBP kinase can be recognized by an antibody against the C-terminal region of a family of p21Cdc42/Rac-activated kinases (PAKs). By using this antibody and a PAK2-specific antibody against the N-terminal region of PAK2 as tools, we further demonstrated that heat shock can induce cleavage of PAK2 to generate a 36-kDa C-terminal catalytic fragment in mouse Balb/c 3T3 and human Hep 3B cells. The kinetic profile of appearance of the 36-kDa C-terminal catalytic fragment of PAK2 matched exactly with the activation of the 36-kDa MBP kinase in these cells induced by heat shock. In addition, the heat shock-induced cleavage and activation of PAK2 was found to be closely associated with both DNA fragmentation and activation of an ICE/CED-3 family cysteine protease termed caspase-3 in heat shock-treated Hep 3B cells. Moreover, blockage of the activation of caspase-3 by pretreating the cells with two specific tetrapeptidic inhibitors of caspases (Ac-DEVD-cho and Ac-YVAD-cmk) could substantially diminish the extent of heat shock-induced cleavage/activation of PAK2. Overall, our results point out that PAK2 is cleaved and activated during the heat shock-induced apoptotic cell death process and suggest that caspase-3 is involved in this process.  相似文献   

14.
The beta-xylosidase from Aspergillus awamori X-100 belonging to the family 3 glycoside hydrolase revealed a distinctive transglycosylating ability to produce xylooligosaccharides with degree of polymerization more than 7. In order to explain this fact, the enzyme has been subjected to the detailed biochemical study. The enzymatic hydrolysis of p-nitrophenyl beta-D-xylopyranoside was found to occur with overall retention of substrate anomeric configuration suggesting cleavage of xylosidic bonds through a double-displacement mechanism. Kinetic study with aryl beta-xylopyranosides substrates, in which leaving group pK(a)s were in the range of 3.96-10.32, revealed monotonic function of log(k(cat)) and no correlation of log(k(cat)/Km) versus pKa values indicating deglycosylation as a rate-limiting step for the enzymatic hydrolysis. The classical bell-shaped pH dependence of k(cat)/Km indicated two ionizable groups in the beta-xylosidase active site with apparent pKa values of 2.2 and 6.4. The kinetic parameters of hydrolysis, Km and k(cat), of p-nitrophenyl beta-D-1,4-xylooligosaccharides were very close to those for hydrolysis of p-nitrophenyl-beta-D-xylopyranoside. Increase of p-nitrophenyl-beta-D-xylopyranoside concentration up to 80 mM led to increasing of the reaction velocity resulting in k(cat)(app)=81 s(-1). Addition of alpha-methyl D-xylopyranoside to the reaction mixture at high concentration of p-nitrophenyl-beta-D-xylopyranoside (50 mM) caused an acceleration of the beta-xylosidase-catalyzed reactions and appearance of a new transglycosylation product, alpha-methyl D-xylopyranosyl-1,4-beta-D-xylopyranoside, that was identified by 1H NMR spectroscopy. The kinetic model suggested for the enzymatic reaction was consistent with the results obtained.  相似文献   

15.
Wang ZX  Zhou B  Wang QM  Zhang ZY 《Biochemistry》2002,41(24):7849-7857
The activities of many protein kinases are regulated by phosphorylation. The phosphorylated protein kinases thus represent an important class of substrates for protein phosphatases. However, our ability to study the phosphatase-catalyzed substrate dephosphorylation has been limited in many cases by the difficulty in preparing sufficient amount of stoichiometrically phosphorylated kinases. We have applied the kinetic theory of substrate reaction during irreversible modification of enzyme activity to the study of phosphatase-catalyzed regulation of kinase activity. As an example, we measured the effect of the hematopoietic protein-tyrosine phosphatase (HePTP) on the reaction catalyzed by the fully activated, bisphosphorylated extracellular signal-regulated protein kinase 2 (ERK2/pTpY). Because only a catalytic amount of ERK2/pTpY is required, this method alleviates the need for large quantities of phospho-ERK2. Kinetic analysis of the ERK2/pTpY-catalyzed substrate reaction in the presence of HePTP leads to the determination of the rate constants for the HePTP-catalyzed dephosphorylation of free ERK2/pTpY and ERK2/pTpY*substrate(s) complexes. The data indicate that ERK2/pTpY is a highly efficient substrate for HePTP (k(cat)/K(m) = 3.05 x 10(6) M(-1) s(-1)). The data also show that binding of ATP to ERK2/pTpY has no effect on ERK2/pTpY dephosphorylation by HePTP. In contrast, binding of an Elk-1 peptide substrate to ERK2/pTpY completely blocks the HePTP action. This result indicates that phosphorylation of Tyr185 is important for ERK2 substrate recognition and that binding of the Elk-1 peptide substrate to ERK2/pTpY blocks the accessibility of pTyr185 to HePTP for dephosphorylation. Collectively, the results establish that the kinetic theory of irreversible enzyme modification can be applied to study the phosphatase catalyzed regulation of kinase activity.  相似文献   

16.
The solvent kinetic isotope effects (SKIE) on the yeast alpha-glucosidase-catalyzed hydrolysis of p-nitrophenyl and methyl-d-glucopyranoside were measured at 25 degrees C. With p-nitrophenyl-D-glucopyranoside (pNPG), the dependence of k(cat)/K(m) on pH (pD) revealed an unusually large (for glycohydrolases) solvent isotope effect on the pL-independent second-order rate constant, (DOD)(k(cat)/K(m)), of 1.9 (+/-0.3). The two pK(a)s characterizing the pH profile were increased in D(2)O. The shift in pK(a2) of 0.6 units is typical of acids of comparable acidity (pK(a)=6.5), but the increase in pK(a1) (=5.7) of 0.1 unit in going from H(2)O to D(2)O is unusually small. The initial velocities show substrate inhibition (K(is)/K(m) approximately 200) with a small solvent isotope effect on the inhibition constant [(DOD)K(is)=1.1 (+/-0.2)]. The solvent equilibrium isotope effects on the K(is) for the competitive inhibitors D-glucose and alpha-methyl D-glucoside are somewhat higher [(DOD)K(i)=1.5 (+/-0.1)]. Methyl glucoside is much less reactive than pNPG, with k(cat) 230 times lower and k(cat)/K(m) 5 x 10(4) times lower. The solvent isotope effect on k(cat) for this substrate [=1.11 (+/-0. 02)] is lower than that for pNPG [=1.67 (+/-0.07)], consistent with more extensive proton transfer in the transition state for the deglucosylation step than for the glucosylation step.  相似文献   

17.
Cytosine deaminase (CDA) from Escherichia coli was shown to catalyze the deamination of isoguanine (2-oxoadenine) to xanthine. Isoguanine is an oxidation product of adenine in DNA that is mutagenic to the cell. The isoguanine deaminase activity in E. coli was partially purified by ammonium sulfate fractionation, gel filtration, and anion exchange chromatography. The active protein was identified by peptide mass fingerprint analysis as cytosine deaminase. The kinetic constants for the deamination of isoguanine at pH 7.7 are as follows: k(cat) = 49 s(-1), K(m) = 72 μM, and k(cat)/K(m) = 6.7 × 10(5) M(-1) s(-1). The kinetic constants for the deamination of cytosine are as follows: k(cat) = 45 s(-1), K(m) = 302 μM, and k(cat)/K(m) = 1.5 × 10(5) M(-1) s(-1). Under these reaction conditions, isoguanine is the better substrate for cytosine deaminase. The three-dimensional structure of CDA was determined with isoguanine in the active site.  相似文献   

18.
J Shaffer  J A Adams 《Biochemistry》1999,38(37):12072-12079
The dissociation rate constants for the two products of the reaction catalyzed by protein kinase A, ADP and phosphopeptide, were measured using a catalytic trapping technique to determine the role of product release in enzyme turnover. The enzyme was preequilibrated with ADP, and the reaction was initiated with a peptide substrate, LRRASLG, and ATP in a rapid quench flow instrument. At high, free magnesium concentrations (>2 mM), the large 'burst' in phosphopeptide production disappears, and, at low concentrations of free magnesium (0.5-1 mM), the kinetic transients become sigmoidal prior to the linear turnover phase. Increasing the concentrations of ATP or ADP did not influence the shape of the kinetic transients in the first 20 ms. ADP preequilibration protects the enzyme from inhibition by the covalent inactivator p-fluorosulfonylbenzoyl 5'-adenosine at 0.5 mM free magnesium, indicating that a competent E. ADP complex forms at low metal concentrations and the sigmoidal behavior in the catalytic trapping experiment is not due to free enzyme at high ATP concentrations. Simulations of the data indicate that ADP release is rate-limiting for turnover at high magnesium concentrations, but, at lower physiological levels of 0.5 and 1 mM, the off rate of ADP is 3- and 2-fold higher than k(cat), respectively. In contrast, the initial portions of the kinetic transients at 0.5 mM free magnesium were unaffected by phosphopeptide preequilibration, indicating that the release rate of this product is significantly larger than turnover. The transient kinetic data, coupled with a previous report [Shaffer and Adams (1999) Biochemistry 38, 5572-5581], support a phosphorylation mechanism under physiological magnesium concentrations that incorporates two partially rate-determining conformational changes, one prior to and one after the phosphoryl transfer step. We propose that the initial step activates the enzyme through key positioning of one or more active-site residues and the second step relaxes this conformation, a prerequisite for a subsequent catalytic cycle.  相似文献   

19.
The pyruvate dehydrogenase (E1) component of the pyruvate dehydrogenase complex (PDC) catalyzes a two-step reaction. Recombinant production of substrate amounts of the lipoyl domains of the dihydrolipoyl transacetylase (E2) component of the mammalian PDC allowed kinetic characterization of the rapid physiological reaction catalyzed by E1. Using either the N-terminal (L1) or the internal (L2) lipoyl domain of E2 as a substrate, analyses of steady state kinetic data support a ping pong mechanism. Using standard E1 preparations, Michaelis constants (Km) were 52 +/- 14 microM for L1 and 24.8 +/- 3.8 microM for pyruvate and k(cat) was 26.3 s(-1). With less common, higher activity preparations of E1, the Km values were > or =160 microM for L1 and > or =35 microM for pyruvate and k(cat) was > or =70 s(-1). Similar results were found with the L2 domain. The best synthetic lipoylated-peptide (L2 residues 163-177) was a much poorer substrate (Km > or =15 mM, k(cat) approximately equals 5 s(-1); k(cat)/Km decreased >1,500-fold) than L1 or L2, but a far better substrate in the E1 reaction than free lipoamide (k(cat)/Km increased >500-fold). Each lipoate source was an effective substrate in the dihydrolipoyl dehydrogenase (E3) reaction, but E3 had a lower Km for the L2 domain than for lipoamide or the lipoylated peptides. In contrast to measurements with slow E1 model reactions that use artificial acceptors, we confirmed that the natural E1 reaction, using lipoyl domain acceptors, was completely inhibited (>99%) by phosphorylation of E1 and the phosphorylation strongly inhibited the reverse of the second step catalyzed by E1. The mechanisms by which phosphorylation interferes with E1 activity is interpreted based on accrued results and the location of phosphorylation sites mapped onto the 3-D structure of related alpha-keto acid dehydrogenases.  相似文献   

20.
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