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 93 毫秒
1.
Ross L Stein 《Biochemistry》2002,41(3):991-1000
Aryl acylamidase (EC 3.1.5.13; AAA) catalyzes the hydrolysis of p-nitroacetanilide (PNAA) via the standard three-step mechanism of serine hydrolases: binding of substrate (K(s)), acylation of active-site serine (k(acyl)), and hydrolytic deacylation (k(deacyl)). Key mechanistic findings that emerged from this study include that (1) AAA requires a deprotonated base with a pK(a) of 8.3 for expression of full activity toward PNAA. Limiting values of kinetic parameters at high pH are k(c) = 7 s(-1), K(m) = 20 microM, and k(c)/K(m) = 340 000 M(-1) s(-1). (2) At pH 10, where all the isotope effects were conducted, k(c) is equally rate-limited by k(acyl) and k(deacyl). (3) The following isotope effects were determined: (D)()2(O)(k(c)/K(m)) = 1.7 +/- 0.2, (D)()2(O)k(c) = 3.5 +/- 0.3, and (beta)(D)(k(c)/K(m)) = 0.83 +/- 0.04, (beta)(D)k(c) = 0.96 +/- 0.01. These values, together with proton inventories for k(c)/K(m) and k(c), suggest the following mechanism: (i) The initial binding of substrate to enzyme to form the Michaelis complex is accompanied by solvation changes that generate solvent deuterium isotope effects originating from hydrogen ion fractionation at multiple sites on the enzyme surface. (ii) From within the Michaelis complex, the active site serine attacks the carbonyl carbon of PNAA with general-base catalysis to form a substantially tetrahedral transition state enroute to the acyl-enzyme. (iii) Finally, deacylation occurs through a process involving a rate-limiting solvent isotope effect, generating conformational change of the acyl-enzyme that positions the carbonyl bond in a polarizing environment that is optimal for attack by water.  相似文献   

2.
Case A  Stein RL 《Biochemistry》2006,45(7):2443-2452
Ubiquitin C-terminal hydrolases (UCHs) cleave Ub-X bonds (Ub is ubiquitin and X an alcohol, an amine, or a protein) through a thioester intermediate that is produced by nucleophilic attack of the Cys residue of a Cys-SH/His-Im catalytic diad. We are studying the mechanism of UCH-L1, a UCH that is implicated in Parkinson's disease, and now wish to report our initial findings. (i) Pre-steady-state kinetic studies for UCH-L1-catalyzed hydrolysis of Ub-AMC (AMC, 7-amido-4-methylcoumarin) indicate that k(cat) is rate-limited by acyl-enzyme formation. Thus, K(m) = K(s), the dissociation constant for the Michaelis complex, and k(cat) = k(2), the rate constant for acyl-enzyme formation. (ii) For K(assoc) (=K(s)(-)(1)), DeltaC(p) = -0.8 kcal mol(-)(1) deg(-)(1) and is consistent with coupling between substrate association and a conformational change of the enzyme. For k(2), DeltaS(++) = 0 and suggests that in the E-S, substrate and active site residues are precisely aligned for reaction. (iii) Solvent isotope effects are (D)K(assoc) = 0.5 and (D)k(2) = 0.9, suggesting that the substrate binds to a form of free enzyme in which the active site Cys exists as the thiol. In the resultant Michaelis complex, the diad has tautomerized to ion pair Cys-S(-)/His-ImH(+). Subsequent attack of thiolate produces the acyl-enzyme species. In contrast, isotope effects for association of UCH-L1 with transition-state analogue ubiquitin aldehyde suggest that an alternative mechanistic pathway can sometimes be available to UCH-L1 involving general base-catalyzed attack of Cys-SH by His-Im.  相似文献   

3.
Chesnokova LS  Witt SN 《Biochemistry》2005,44(33):11224-11233
Hsp70 chaperones are heterotropic allosteric systems in which ATP and misfolded or aggregated polypeptides are the activating ligands. To gain insight into the mechanism by which ATP and polypeptides regulate Hsp70 chaperone activity, the effect of a short peptide on the K(M) for ATP was analyzed using the Escherichia coli Hsp70 called DnaK. In the absence of peptide, the K(-P)(M) for ATP is 52 +/- 11 nM, whereas this value jumps to 14.6 +/- 1.6 microM in the presence of saturating peptide. This finding supports a mechanism in which ATP binding drives the chaperone in one direction and peptide binding pushes the chaperone back in the opposite direction (and thus increases K(M)), according to ATP + DnaK.P <==> ATP.DnaK.P <==> ATP.DnaK* + P, where ATP.DnaK.P is an intermediate from which competing ATP hydrolysis occurs (ATP.DnaK.P --> ADP.DnaK.P). We show that this branched mechanism can even explain how DnaK hydrolyzes ATP in the absence of peptide and that the true rate constant for DnaK-mediated ATP hydrolysis (k(hy)) in the absence of peptide may be as high as 0.5 s(-)(1) (rather than 5 x 10(-)(4) s(-)(1) as often stated in the literature). What happens is that a conformational equilibrium outcompetes ATP hydrolysis and effectively reduces the concentration of the intermediate by a factor of a thousand, resulting in the following relation: k(cat) = k(hy)/1000 = 5 x 10(-)(4) s(-)(1). How polypeptide substrates and the co-chaperone DnaJ modulate DnaK to achieve its theoretical maximal rate of ATP hydrolysis, which we suggest is 0.5 s(-)(1), is discussed.  相似文献   

4.
Hsc66 from Escherichia coli is a constitutively expressed hsp70 class molecular chaperone whose activity is coupled to ATP binding and hydrolysis. To better understand the mechanism and regulation of Hsc66, we investigated the kinetics of ATP hydrolysis and the interactions of Hsc66 with nucleotides. Steady-state experiments revealed that Hsc66 has a low affinity for ATP (K(m)(ATP) = 12.7 microM) compared with other hsp70 chaperones. The kinetics of nucleotide binding were determined by analyzing changes in the Hsc66 absorbance spectrum using stopped-flow methods at 23 degrees C. ATP binding results in a rapid, biphasic increase of Hsc66 absorbance at 280 nm; this is interpreted as arising from a two-step process in which ATP binding (k(a)(ATP) = 4.2 x 10(4) M(-1) s(-1), k(d)(ATP) = 1.1 s(-1)) is followed by a slow conformational change (k(conf) = 0. 1 s(-1)). Under single turnover conditions, the ATP-induced transition decays exponentially with a rate (k(decay) = 0.0013 s(-1)) similar to that observed in both steady-state and single turnover ATP hydrolysis experiments (k(hyd) = 0.0014 s(-1)). ADP binding to Hsc66 results in a monophasic transition in the absence (k(a)(ADP) = 7 x 10(5) M(-1) s(-1), k(d)(ADP) = 60 s(-1)) and presence of physiological levels of inorganic phosphate (k(a)(ADP(P(i)) = 0.28 x 10(5) M(-1) s(-1), k(d)(ADP(P(i)) = 9.1 s(-1)). These results indicate that ATP hydrolysis is the rate-limiting step under steady-state conditions and is >10(3)-fold slower than the rate of ADP/ATP exchange. Thus, in contrast to DnaK and eukaryotic forms of hsp70 that have been characterized to date, the R if T equilibrium balance for Hsc66 is shifted in favor of the low peptide affinity T state, and regulation of the reaction cycle is expected to occur at the ATP hydrolysis step rather than at nucleotide exchange.  相似文献   

5.
Eubanks LM  Poulter CD 《Biochemistry》2003,42(4):1140-1149
1-Deoxy-d-xylulose 5-phosphate synthase (DXP synthase) catalyzes the thiamine diphosphate (TPP)-dependent condensation of pyruvate and d-glyceraldehyde 3-phosphate (GAP) to yield DXP in the first step of the methylerythritol phosphate pathway for isoprenoid biosynthesis. Steady-state kinetic constants for DXP synthase calculated from the initial velocities measured at varying concentrations of substrates were as follows: k(cat) = 1.9 +/- 0.1 s(-1), K(m)(GAP) = 0.068 +/- 0.001 mM, and K(m)(pyruvate) = 0.44 +/- 0.05 mM for pyruvate and GAP; k(cat) = 1.7 +/- 0.1 s(-1), K(m)(d-glyceraldehyde) = 33 +/- 3 mM, and K(m)(pyruvate) = 1.9 +/- 0.5 mM for d-glyceraldehyde and pyruvate. beta-Fluoropyruvate was investigated as a dead-end inhibitor for pyruvate. Double-reciprocal plots showed a competitive inhibition pattern with respect to pyruvate and noncompetitive inhibition with respect to GAP/d-glyceraldehyde. (14)CO(2) trapping experiments demonstrated that the binding of both substrates (pyruvate and GAP/d-glyceraldehyde) is required for the formation of a catalytically competent enzyme-substrate complex. These results are consistent with an ordered mechanism for DXP synthase where pyruvate binds before GAP/d-glyceraldehyde.  相似文献   

6.
To understand the molecular mechanism underlying the diversity of mammalian skeletal muscle fibers, the elementary steps of the cross-bridge cycle were investigated in three fast-twitch fiber types from rabbit limb muscles. Skinned fibers were maximally Ca(2+)-activated at 20 degrees C and the effects of MgATP, phosphate (P, P(i)), and MgADP were studied on three exponential processes by sinusoidal analysis. The fiber types (IIA, IID, and IIB) were determined by analyzing the myosin heavy-chain isoforms after mechanical experiments using high-resolution SDS-PAGE. The results were consistent with the following cross-bridge scheme: where A is actin, M is myosin, D is MgADP, and S is MgATP. All states except for those in brackets are strongly bound states. All rate constants of elementary steps (k(2), 198-526 s(-1); k(-2), 51-328 s(-1); k(4), 13.6-143 s(-1); k(-4), 13.6-81 s(-1)) were progressively larger in the order of type IIA, type IID, and type IIB fibers. The rate constants of a transition from a weakly bound state to a strongly bound state (k(-2), k(4)) varied more among fiber types than their reversals (k(2), k(-4)). The equilibrium constants K(1) (MgATP affinity) and K(2) (=k(2)/k(-2), ATP isomerization) were progressively less in the order IIA, IID, and IIB. K(4) (=k(4)/k(-4), force generation) and K(5) (P(i) affinity) were larger in IIB than IIA and IID fibers. K(1) showed the largest variation indicating that the myosin head binds MgATP more tightly in the order IIA (8.7 mM(-1)), IID (4.9 mM(-1)), and IIB (0.84 mM(-1)). Similarly, the MgADP affinity (K(0)) was larger in type IID fibers than in type IIB fibers.  相似文献   

7.
An effective means of relieving the toxicity of furan aldehydes, furfural (FFA) and 5-hydroxymethylfurfural (HMF), on fermenting organisms is essential for achieving efficient fermentation of lignocellulosic biomass to ethanol and other products. Ari1p, an aldehyde reductase from Saccharomyces cerevisiae, has been shown to mitigate the toxicity of FFA and HMF by catalyzing the NADPH-dependent conversion to corresponding alcohols, furfuryl alcohol (FFOH) and 5-hydroxymethylfurfuryl alcohol (HMFOH). At pH 7.0 and 25°C, purified Ari1p catalyzes the NADPH-dependent reduction of substrates with the following values (k(cat) (s(-1)), k(cat)/K(m) (s(-1)mM(-1)), K(m) (mM)): FFA (23.3, 1.82, 12.8), HMF (4.08, 0.173, 23.6), and dl-glyceraldehyde (2.40, 0.0650, 37.0). When acting on HMF and dl-glyceraldehyde, the enzyme operates through an equilibrium ordered kinetic mechanism. In the physiological direction of the reaction, NADPH binds first and NADP(+) dissociates from the enzyme last, demonstrated by k(cat) of HMF and dl-glyceraldehyde that are independent of [NADPH] and (K(ia)(NADPH)/k(cat)) that extrapolate to zero at saturating HMF or dl-glyceraldehyde concentration. Microscopic kinetic parameters were determined for the HMF reaction (HMF+NADPH?HMFOH+NADP(+)), by applying steady-state, presteady-state, kinetic isotope effects, and dynamic modeling methods. Release of products, HMFOH and NADP(+), is 84% rate limiting to k(cat) in the forward direction. Equilibrium constants, [NADP(+)][FFOH]/[NADPH][FFA][H(+)]=5600×10(7)M(-1) and [NADP(+)][HMFOH]/[NADPH][HMF][H(+)]=4200×10(7)M(-1), favor the physiological direction mirrored by the slowness of hydride transfer in the non-physiological direction, NADP(+)-dependent oxidation of alcohols (k(cat) (s(-1)), k(cat)/K(m) (s(-1)mM(-1)), K(m) (mM)): FFOH (0.221, 0.00158, 140) and HMFOH (0.0105, 0.000104, 101).  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.
Taly A  Baciou L  Sebban P 《FEBS letters》2002,532(1-2):91-96
Photosynthetic reaction centers (RCs) from Rhodobacter sphaeroides were incorporated in dimyristoylphosphatidylcholine (DMPC) liposomes. The first and second electron transfer rates (k(AB)(1) and k(AB)(2), respectively) between the first and the second quinone electron acceptors have been measured as a function of temperature, across the phase transition of DMPC (23 degrees C). The Eyring plots of k(AB)(1) display straight lines. In contrast, the Eyring plots for k(AB)(2) in proteoliposomes show a break at about 23.5 degrees C. This physical discrimination between the two electron transfer reactions demonstrates that the stiffness of the lipid environment of the RCs and/or the protein-protein interactions influence the parameters governing k(AB)(2), but not the gating process limiting k(AB)(1).  相似文献   

10.
To elucidate a detailed catalytic mechanism for nitrile hydratases (NHases), the pH and temperature dependence of the kinetic constants k(cat) and K(m) for the cobalt-type NHase from Pseudonocardia thermophila JCM 3095 (PtNHase) were examined. PtNHase was found to exhibit a bell-shaped curve for plots of relative activity versus pH at pH 3.2-11 and was found to display maximal activity between pH 7.2 and 7.8. Fits of these data provided pK(E)(S1) and pK(E)(S2) values of 5.9 +/- 0.1 and 9.2 +/- 0.1 (k(cat)' = 130 +/- 1 s(-1)), respectively, and pK(E)(1) and pK(E)(2) values of 5.8 +/- 0.1 and 9.1 +/- 0.1 (k(cat)'/K(m)' = (6.5 +/- 0.1) x 10(3) s(-1) mm(-1)), respectively. Proton inventory studies indicated that two protons are transferred in the rate-limiting step of the reaction at pH 7.6. Because PtNHase is stable at 60 degrees C, an Arrhenius plot was constructed by plotting ln(k(cat)) versus 1/T, providing E(a) = 23.0 +/- 1.2 kJ/mol. The thermal stability of PtNHase also allowed DeltaH(0) ionization values to be determined, thus helping to identify the ionizing groups exhibiting the pK(E)(S1) and pK(E)(S2) values. Based on DeltaH(0)(ion) data, pK(E)(S1) is assigned to betaTyr(68), whereas pK(E)(S2) is assigned to betaArg(52), betaArg(157), or alphaSer(112) (NHases are alpha(2)beta(2)-heterotetramers). A combination of these data with those previously reported for NHases and synthetic model complexes, along with sequence comparisons of both iron- and cobalt-type NHases, allowed a novel catalytic mechanism for NHases to be proposed.  相似文献   

11.
Stopped-flow measurements have been employed to study the kinetics of the conformational changes in TetR (B) induced by tetracycline binding with and without Mg(2+) ions. Result of stopped-flow fluorometry measurements at pH 8.0 indicate conformational changes in the helix-turn-helix motif in the N-terminal domain and in the C-terminal inducer binding domain. Binding of tetracycline (Tc) to TetR in the absence of Mg(2+) can be described by a simple kinetics process, which is limited to the first step association without any unimolecular conformational change step upon Tc binding. The rate constants for this process are equal to 2.0 x 10(5) M(-)(1) s(-)(1) and 2.1 s(-)(1) for the forward and backward reaction, respectively, and gave the binding constant K(a) = 0.96 x 10(5) M(-)(1). The kinetics of [Tc-Mg](+) binding to TetR can be described by reactions in which the first step describes the association characterized by the rate constants k(a) = 1.4 x 10(5) M(-)(1) s(-)(1) and k(d) = 2.2 x 10(-)(2) s(-)(1) and binding constant K(a) = 6.3 x 10(6) M(-)(1). The first step of [Tc-Mg](+) association is followed by at least three conformational change steps, which occur in the inducer binding site and then propagate to the surroundings of Trp75 and Trp43 residues. The rate constants for the forward, k(c), and backward, k(-)(c), reaction for each of these conformational steps have been determined. The thermodynamics of the binding of tetracycline with and without Mg(2+) to TetR was investigated by isothermal titration calorimetry (ITC) at pH 8.0 and 25 degrees C. The measurement shows that TetR dimer possesses two equivalent binding sites for tetracycline, characterized by binding constant K(a) = 9.0 x 10(6) M(-)(1) and K(a) = 7.0 x 10(4) M(-)(1) for Tc with and without Mg(2+), respectively. The binding of the inducer to TetR, in the presence and absence of Mg(2+) ion, is an enthalpy-driven reaction characterized by DeltaH = -51 kJ mol(-)(1) and DeltaH = -33 kJ mol(-)(1), respectively. The entropy change, DeltaS, for the interaction in the presence of Mg(2+) is equal to -38.9 J K(-)(1) mol(-)(1), and for the tetracycline alone, it was estimated at -17.6 J K(-)(1) mol(-)(1).  相似文献   

12.
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   

13.
Human angiotensin-converting enzyme-related carboxypeptidase (ACE2) is a zinc metalloprotease whose closest homolog is angiotensin I-converting enzyme. To begin to elucidate the physiological role of ACE2, ACE2 was purified, and its catalytic activity was characterized. ACE2 proteolytic activity has a pH optimum of 6.5 and is enhanced by monovalent anions, which is consistent with the activity of ACE. ACE2 activity is increased approximately 10-fold by Cl(-) and F(-) but is unaffected by Br(-). ACE2 was screened for hydrolytic activity against a panel of 126 biological peptides, using liquid chromatography-mass spectrometry detection. Eleven of the peptides were hydrolyzed by ACE2, and in each case, the proteolytic activity resulted in removal of the C-terminal residue only. ACE2 hydrolyzes three of the peptides with high catalytic efficiency: angiotensin II () (k(cat)/K(m) = 1.9 x 10(6) m(-1) s(-1)), apelin-13 (k(cat)/K(m) = 2.1 x 10(6) m(-1) s(-1)), and dynorphin A 1-13 (k(cat)/K(m) = 3.1 x 10(6) m(-1) s(-1)). The ACE2 catalytic efficiency is 400-fold higher with angiotensin II () as a substrate than with angiotensin I (). ACE2 also efficiently hydrolyzes des-Arg(9)-bradykinin (k(cat)/K(m) = 1.3 x 10(5) m(-1) s(-1)), but it does not hydrolyze bradykinin. An alignment of the ACE2 peptide substrates reveals a consensus sequence of: Pro-X((1-3 residues))-Pro-Hydrophobic, where hydrolysis occurs between proline and the hydrophobic amino acid.  相似文献   

14.
N-benzoyl-L-phenylalanyl-L-phenylalanine is an excellent peptide substrate for carboxy-peptidase A; at 30 degrees C and pH 7.5, K(m) is 2.6 x 10(-5) M while k(cat) is 177 s(-1) (k(cat)/K(m) = 6.8 x 10(6) M(-1) s(-1)). Indole-3-acetic acid is a noncompetitive or mixed inhibitor towards the peptide and toward hippuryl-L-phenylalanine; plots of E/V vs [Inhibitor] are linear. N-Benzoyl-L-phenylalanine is a competitive inhibitor of peptide hydrolysis, and plots of E/V vs [Inhibitor] are again linear. One molecule of inhibitor binds per active site, and these inhibitors bind in different sites. At constant peptide substrate concentration and a series of constant concentrations of indole-3-acetic acid, plots of E/V vs the concentration of N-benzoyl-L-phenylalanine are linear and intersect behind the E/V axis and above the [Inhibitor] axis. This shows that both inhibitors can bind simultaneously and that binding of one facilitates the binding of the other (beta = 0.18). Employing the ester substrate hippuryl-DL,beta-phenyllactate, the same type of behavior is observed in the reverse sense; N-benzoyl-L-phenylalanine is a linear noncompetitive inhibitor and indole-3-acetic acid is a linear competitive inhibitor. Again the two inhibitor plot is linear and intersects above the [Inhibitor] axis (beta = 0.12). Previous X-ray crystallographic studies have indicated that indole-3-acetic acid binds in the hydrophobic pocket of the S'(1) site, while N-benzoyl-L-phenylalanine binds in the S(1)-S(2) site. The product complex for hydrolysis of N-benzoyl-L-phenylalanyl-L-phenylalanine (phenylalanine + N-benzoyl-L-phenylalanine) occupies both of these sites. However, the present work shows that the peptide substrate does not bind to the enzyme at pH 7.5 so as to be competitive with indole-3-acetic acid. The binding sites may be formed via conformational changes induced or stabilized by substrate and product binding. Copyright 2000 Academic Press.  相似文献   

15.
Stein RL  DeCicco C  Nelson D  Thomas B 《Biochemistry》2001,40(19):5804-5811
gamma-Glutamyl transpeptidase (gammaGTase) catalyzes the transfer of the gamma-glutamyl moiety of gamma-glutamyl-derived peptides, such as glutathione (gammaGlu-Cys-Gly), and anilides, such as gamma-glutamyl-7-amido-4-methylcoumarin (gammaGlu-AMC), to acceptor molecules, including water and various dipeptides. These acyl-transfer reactions all occur through a common acyl-enzyme intermediate formed from attack of an active site hydroxyl on the gamma-carbonyl carbon of gammaGlu-X with displacement of X. In this paper, we report that gammaGTase is potently inhibited by the gamma-boronic acid analogue of L-glutamic acid, 3-amino-3-carboxypropaneboronic acid (gamma-boroGlu). We propose that gamma-boroGlu adds to the active site hydroxyl of gammaGTase to form a covalent, tetrahedral adduct that resembles tetrahedral transition states and intermediates that occur along the reaction pathway for gammaGTase-catalyzed reactions. Our studies demonstrate that gamma-boroGlu is a competitive inhibitor of the gammaGTase-catalyzed hydrolysis of gammaGlu-AMC with a K(i) value of 35 nM. Kinetics of inhibition studies allow us to estimate the following values: k(on) = 400 mM(-1) s(-1) and k(off) = 0.02 s(-1). We also found that gamma-boroGlu is an uncompetitive inhibitor of Gly-Gly-promoted transamidation of gammaGlu-AMC. This observation is consistent with the kinetic mechanism we determined for gammaGTase-catalyzed transamidation of gammaGlu-AMC by Gly-Gly to form gammaGlu-Gly-Gly. To probe rate-limiting transition states for gammaGTase catalysis and inhibition, we determined solvent deuterium isotope effects. Solvent isotope effects on k(c)/K(m) for hydrolysis of gammaGlu-AMC and k(on) for inhibition by gamma-boroGlu are identical and equal unity, suggesting that the processes governed by these rate constants are both rate-limited by a step that is insensitive to solvent deuterium such as a conformational fluctuation of the initially formed E-S or E-I complex. In contrast, the solvent isotope effect on k(c) is 2.4. k(c) is rate-limited by hydrolysis of the acyl-enzyme intermediate that is formed during reaction of gammaGTase with gammaGlu-AMC. Thus, the magnitude of this isotope effect suggests the formation of a catalytically important protonic bridge in the rate-limiting transition state for deacylation.  相似文献   

16.
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   

17.
Kinetic and binding studies involving a model DNA cytosine-5-methyltransferase, M.HhaI, and a 37-mer DNA duplex containing a single hemimethylated target site were applied to characterize intermediates on the reaction pathway. Stopped-flow fluorescence studies reveal that cofactor S-adenosyl-l-methionine (AdoMet) and product S-adenosyl-l-homocysteine (AdoHcy) form similar rapidly reversible binary complexes with the enzyme in solution. The M.HhaI.AdoMet complex (k(off) = 22 s(-)1, K(D) = 6 microm) is partially converted into products during isotope-partitioning experiments, suggesting that it is catalytically competent. Chemical formation of the product M.HhaI.(Me)DNA.AdoHcy (k(chem) = 0.26 s(-)1) is followed by a slower decay step (k(off) = 0.045 s(-)1), which is the rate-limiting step in the catalytic cycle (k(cat) = 0.04 s(-)1). Analysis of reaction products shows that the hemimethylated substrate undergoes complete (>95%) conversion into fully methylated product during the initial burst phase, indicating that M.HhaI exerts high binding selectivity toward the target strand. The T250N, T250D, and T250H mutations, which introduce moderate perturbation in the catalytic site, lead to substantially increased K(D)(DNA(ternary)), k(off)(DNA(ternary)), K(M)(AdoMet(ternary)) values but small changes in K(D)(DNA(binary)), K(D)(AdoMet(binary)), k(chem), and k(cat). When the target cytosine is replaced with 5-fluorocytosine, the chemistry step leading to an irreversible covalent M.HhaI.DNA complex is inhibited 400-fold (k(chem)(5FC) = 0.7 x 10(-)3 s(-)1), and the Thr-250 mutations confer further dramatic decrease of the rate of the covalent methylation k(chem). We suggest that activation of the pyrimidine ring via covalent addition at C-6 is a major contributor to the rate of the chemistry step (k(chem)) in the case of cytosine but not 5-fluorocytosine. In contrast to previous reports, our results imply a random substrate binding order mechanism for M.HhaI.  相似文献   

18.
Fox DT  Poulter CD 《Biochemistry》2005,44(23):8360-8368
The mechanism of the reaction catalyzed by 2-C-methyl-d-erythritol 4-phosphate (MEP) synthase from Escherichia coli has been studied by steady-state and single-turnover kinetic experiments for the 1-deoxy-d-xylulose 5-phosphoric acid (DXP) analogues, 1,1,1-trifluoro-1-deoxy-d-xylulose 5-phosphoric acid (CF(3)-DXP), 1,1-difluoro-1-deoxy-d-xylulose 5-phosphoric acid (CF(2)-DXP), 1-fluoro-1-deoxy-d-xylulose 5-phosphoric acid (CF-DXP), and 1,2-dideoxy-d-hexulose 6-phosphate (Et-DXP). CF(3)-DXP, CF(2)-DXP, and Et-DXP were poor inhibitors, most likely because of the increase in steric bulk at C1 of DXP. The three analogues were also poor substrates for the enzyme. In contrast, CF-DXP was a good substrate (k(cat)(CF)(-)(DXP) = 37 +/- 2 s(-)(1), K(m)(CF)(-)(DXP) = 227 +/- 25 microM) for MEP synthase when compared to DXP (k(cat)(DXP) = 29 +/- 1 s(-)(1), K(m)(DXP) = 45 +/- 4 microM). A primary deuterium isotope effect was observed under single-turnover conditions when CF-DXP was incubated with 4S-[(2)H]NADPH ((H)k/(D)k = 1.34 +/-0.01), whereas no isotope effect was observed upon incubation with DXP and 4S-[(2)H]NADPH ((H)k/(D)k = 1.02 +/- 0.02). The reaction did not exhibit burst kinetics for either substrate, indicating that product release is not rate-limiting. These studies suggest that positive charge does not develop at C2 of DXP during catalysis. In addition, the isotope effect with CF-DXP and 4S-[(2)H]NADPH but not DXP indicates that the rearrangement step, which precedes hydride transfer, is rate-limiting for DXP but becomes partially rate-limiting for CF-DXP. Thus, rearrangement appears to be enhanced by substitution of a hydrogen atom in the methyl group of DXP by fluorine. These observations are consistent with a retro-aldol/aldol mechanism for the rearrangement during conversion of DXP to MEP.  相似文献   

19.
Rao S  Holz RC 《Biochemistry》2008,47(46):12057-12064
In order to gain insight into the catalytic mechanism of Fe-type nitrile hydratases (NHase), the pH and temperature dependence of the kinetic parameters k cat, K m, and k cat/ K m along with the solvent isotope effect were examined for the Fe-type NHase from Comamonas testosteroni Ni1 ( CtNHase). CtNHase was found to exhibit a bell-shaped curve for plots of relative activity vs pH over pH values 4-10 for the hydration of acrylonitrile and was found to display maximal activity at pH approximately 7.2. Fits of these data provided a p K ES1 value of 6.1 +/- 0.1, a p K ES2 value of 9.1 +/- 0.2 ( k' cat = 10.1 +/- 0.3 s (-1)), a p K E1 value of 6.2 +/- 0.1, and a p K E2 value of 9.2 +/- 0.1 ( k' cat/ K' m of 2.0 +/- 0.2 s (-1) mM (-1)). Proton inventory studies indicate that two protons are transferred in the rate-limiting step of the reaction at pH 7.2. Since CtNHase is stable to 25 degrees C, an Arrhenius plot was constructed by plotting ln( k cat) vs 1/ T, providing an E a of 33.3 +/- 1.5 kJ/mol. Delta H degrees of ionization values were also determined, thus helping to identify the ionizing groups exhibiting the p K ES1 and p K ES2 values. Based on Delta H degrees ion data, p K ES1 is assigned to betaTyr68 while p K ES2 is assigned to betaArg52, betaArg157, or alphaSer116 (NHases are alpha 2beta 2 heterotetramers). Given the strong similarities in the kinetic data obtained for both Co- and Fe-type NHase enzymes, both types of NHase enzymes likely hydrate nitriles in a similar fashion.  相似文献   

20.
Patton SM  Cropp TA  Reynolds KA 《Biochemistry》2000,39(25):7595-7604
The side chain of the antifungal polyketide ansatrienin A produced by Streptomyces collinus contains a cyclohexanecarboxylic acid (CHC) derived moiety. This CHC in the coenzyme A activated form (CHC-CoA) is derived from shikimic acid via a pathway in which the penultimate step is the isomerization of 2-cyclohexenylcarbonyl-CoA to 1-cyclohexenylcarbonyl-CoA. We have purified a 28 kDa 2-cyclohexenylcarbonyl-CoA isomerase (ChcB) from S. collinus and cloned and sequenced the corresponding chcB gene. The predicted amino acid sequence of ChcB showed moderate sequence identity to members of the hydratase/isomerase superfamily of enzymes. The recombinant ChcB was overexpressed in Escherichia coli and purified to homogeneity using metal chelate chromatography. Kinetic analysis demonstrated that recombinant ChcB had wide substrate specificity and could catalyze a double bond isomerization using 2-cyclohexenylcarbonyl-CoA (K(m) 116 +/- 68 microM, k(cat)( )()3.7 +/- 1.0 min(-)(1)), trans-3-hexenyl-CoA (K(m) 39 +/- 10 microM, k(cat)( )()12.8 +/- 1 min(-)(1)), and vinylacetyl-CoA (K(m) 156 +/- 34 microM, k(cat)( )()29 +/- 3 min(-)(1)) as substrates. ChcB activity in cell extracts of S. collinus SP1, an insertionally disrupted chcB mutant, was shown to decrease by more than 99% (as compared to the wild-type strain) using all three of these substrates. The S. collinus SP1 strain, unlike the wild-type strain, could not produce omega-cyclohexyl fatty acids but was still able to grow efficiently on methyl oleate as a sole carbon source. These observations demonstrate that the S. collinus ChcB is required for catalyzing the isomerization of 2-cyclohexenylcarbonyl-CoA to 1-cyclohexenylcarbonyl-CoA during CHC-CoA biosynthesis but not for degradation of unsaturated fatty acids. The chcB gene does not appear to be associated with the ansatrienin biosynthetic gene cluster, which has previously been shown to contain at least one gene known to be essential for CHC-CoA biosynthesis. This finding represents a notable exception to the general rule regarding the clustering of polyketide biosynthetic pathway genes.  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