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.
Cathepsin C, or dipeptidyl peptidase I, is a lysosomal cysteine protease of the papain family that catalyzes the sequential removal of dipeptides from the free N-termini of proteins and peptides. Using the dipeptide substrate Ser-Tyr-AMC, cathepsin C was characterized in both steady-state and pre-steady-state kinetic modes. The pH(D) rate profiles for both log k cat/ K m and log k cat conformed to bell-shaped curves for which an inverse solvent kinetic isotope effect (sKIE) of 0.71 +/- 0.14 for (D)( k cat/ K a) and a normal sKIE of 2.76 +/- 0.03 for (D) k cat were obtained. Pre-steady-state kinetics exhibited a single-exponential burst of AMC formation in which the maximal acylation rate ( k ac = 397 +/- 5 s (-1)) was found to be nearly 30-fold greater than the rate-limiting deacylation rate ( k dac = 13.95 +/- 0.013 s (-1)) and turnover number ( k cat = 13.92 +/- 0.001 s (-1)). Analysis of pre-steady-state burst kinetics in D 2O allowed abstraction of a normal sKIE for the acylation half-reaction that was not observed in steady-state kinetics. Since normal sKIEs were obtained for all measurable acylation steps in the presteady state [ (D) k ac = 1.31 +/- 0.04, and the transient kinetic isotope effect at time zero (tKIE (0)) = 2.3 +/- 0.2], the kinetic step(s) contributing to the inverse sKIE of (D)( k cat/ K a) must occur more rapidly than the experimental time frame of the transient kinetics. Results are consistent with a chemical mechanism in which acylation occurs via a two-step process: the thiolate form of Cys-234, which is enriched in D 2O and gives rise to the inverse value of (D)( k cat/ K a), attacks the substrate to form a tetrahedral intermediate that proceeds to form an acyl-enzyme intermediate during a proton transfer step expressing a normal sKIE. The subsequent deacylation half-reaction is rate-limiting, with proton transfers exhibiting normal sKIEs. Through derivation of 12 equations describing all kinetic parameters and sKIEs for the proposed cathepsin C mechanism, integration of both steady-state and pre-steady-state kinetics with sKIEs allowed the provision of at least one self-consistent set of values for all 13 rate constants in this cysteine protease's chemical mechanism. Simulation of the resulting kinetic profile showed that at steady state approximately 80% of the enzyme exists in an active-site cysteine-acylated form in the mechanistic pathway. The chemical and kinetic details deduced from this work provide a potential roadmap to help steer drug discovery efforts for this and other disease-relevant cysteine proteases.  相似文献   

2.
Bjornson KP  Allen DJ  Modrich P 《Biochemistry》2000,39(11):3176-3183
Escherichia coli MutS protein, which is required for mismatch repair, has a slow ATPase activity that obeys Michalelis-Menten kinetics. At 37 degrees C, the steady-state turnover rate for ATP hydrolysis is 1.0 +/- 0.3 min(-1) per monomer equivalent with a K(m) of 33 +/- 6 microM. Hydrolysis is competitively inhibited by the ATP analogues AMPPNP and ATPgammaS, with K(i) values of 4 microM in both cases, and by ADP with a K(i) of 40 microM. The rate of ATP hydrolysis is stimulated 2-5-fold by short hetero- and homoduplex DNAs. The concentration of DNA cofactor that yields half-maximal stimulation is lowest for oligodeoxynucleotide duplexes that contain a mismatched base pair. Pre-steady-state chemical quench analysis has demonstrated a substoichiometric initial burst of ADP formation by free MutS that is governed by a rate constant of 78 min(-1), indicating that the rate-limiting step for the steady-state reaction occurs after hydrolysis. Prebinding of MutS to homoduplex DNA does not alter the burst kinetics or amplitude but only increases the steady-state rate. In contrast, binding of the protein to heteroduplex DNA abolishes the burst of ADP formation, indicating that the rate-limiting step now occurs before hydrolysis. Gel filtration analysis indicates that the MutS dimer assembles into higher order oligomers in a concentration-dependent manner, and that ATP binding shifts this equilibrium to favor assembly. These results, together with kinetic findings, indicate nonequivalence of subunits within a MutS oligomer with respect to ATP hydrolysis and DNA binding.  相似文献   

3.
The complete time course of the hydrolysis of p-nitrophenyl phosphate catalyzed by the low molecular weight (acid) phosphotyrosyl protein phosphatase from bovine heart was elucidated and analyzed in detail. Burst titration kinetics were demonstrated for the first time with this class of enzyme. At pH 7.0, 4.5 degrees C, a transient pre-steady-state "burst" of p-nitrophenol was formed with a rate constant of 48 s-1. The burst was effectively stoichiometric and corresponded to a single enzyme active site/molecule. The burst was followed by a slow steady-state turnover of the phosphoenzyme intermediate with a rate constant of 1.2 s-1. Product inhibition studies indicated an ordered uni-bi kinetic scheme for the hydrolysis. Partition experiments conducted for several substrates revealed a constant product ratio. Vmax was constant for these substrates, and the overall rate of hydrolysis was increased greatly in the presence of alcohol acceptors. An enzyme-catalyzed 18O exchange between inorganic phosphate and water was detected and occurred with kcat = 4.47 x 10(-3) s-1 at pH 5.0, 37 degrees C. These results were all consistent with the existence of a phosphoenzyme intermediate in the catalytic pathway and with the breakdown of the intermediate being the rate-limiting step. The true Michaelis binding constant Ks = 6.0 mM, the apparent Km = 0.38 mM, and the rate constants for phosphorylation (k2 = 540 s-1) and dephosphorylation (k3 = 36.5 s-1) were determined under steady-state conditions with p-nitrophenyl phosphate at pH 5.0 and 37 degrees C in the presence of phosphate acceptors. The energies of activation for the enzyme-catalyzed hydrolysis at pH 5.0 and 7.0 were 13.6 and 14.1 kcal/mol, respectively. The activation energy for the enzyme-catalyzed medium 18O exchange between phosphate and water was 20.2 kcal/mol. Using the available equilibrium and rate constants, an energetic diagram was constructed for the enzyme-catalyzed reaction.  相似文献   

4.
Ruzzini AC  Horsman GP  Eltis LD 《Biochemistry》2012,51(29):5831-5840
meta-Cleavage product (MCP) hydrolases catalyze C-C bond fission in the aerobic catabolism of aromatic compounds by bacteria. These enzymes utilize a Ser-His-Asp triad to catalyze hydrolysis via an acyl-enzyme intermediate. BphD, which catalyzes the hydrolysis of 2-hydroxy-6-oxo-6-phenylhexa-2,4-dienoic acid (HOPDA) in biphenyl degradation, catalyzed the hydrolysis of an ester analogue, p-nitrophenyl benzoate (pNPB), with a k(cat) value (6.3 ± 0.5 s(-1)) similar to that of HOPDA (6.5 ± 0.5 s(-1)). Consistent with the breakdown of a shared intermediate, product analyses revealed that BphD catalyzed the methanolysis of both HOPDA and pNPB, partitioning the products to benzoic acid and methyl benzoate in similar ratios. Turnover of HOPDA was accelerated up to 4-fold in the presence of short, primary alcohols (methanol > ethanol > n-propanol), suggesting that deacylation is rate-limiting during catalysis. In the steady-state hydrolysis of HOPDA, k(cat)/K(m) values were independent of methanol concentration, while both k(cat) and K(m) values increased with methanol concentration. This result was consistent with a simple model of nucleophilic catalysis. Although the enzyme could not be saturated with pNPB at methanol concentrations of >250 mM, k(obs) values from the steady-state turnover of pNPB at low methanol concentrations were also consistent with a nucleophilic mechanism of catalysis. Finally, transient-state kinetic analysis of pNPB hydrolysis by BphD variants established that substitution of the catalytic His reduced the rate of acylation by more than 3 orders of magnitude. This suggests that for pNPB hydrolysis, the serine nucleophile is activated by the His-Asp dyad. In contrast, rapid acylation of the H265Q variant during C-C bond cleavage suggests that the serinate forms via a substrate-assisted mechanism. Overall, the data indicate that ester hydrolysis proceeds via the same acyl-enzyme intermediate as that of the physiological substrate but that the serine nucleophile is activated via a different mechanism.  相似文献   

5.
Kinetic constants for the hydrolysis by porcine tissue beta-kallikrein B and by bovine trypsin of a number of peptides related to the sequence of kininogen (also one containing a P2 glycine residue instead of phenylalanine) and of a series of corresponding arginyl peptide esters with various apolar P2 residues have been determined under strictly comparative conditions. kcat and kcat/Km values for the hydrolysis of the Arg-Ser bonds of the peptides by trypsin are conspicuously high. kcat for the best of the peptide substrates, Ac-Phe-Arg-Ser-Val-NH2, even reaches kcat for the corresponding methyl ester, indicating rate-limiting deacylation also in the hydrolysis of a peptide bond by this enzyme. kcat/Km for the hydrolysis of the peptide esters with different nonpolar L-amino acids in P2 is remarkably constant (range 1.7), as it is for the pair of the above pentapeptides with P2 glycine or phenylalanine. kcat for the ester substrates varies fivefold, however, being greatest for the P2 glycine compounds. Obviously, an increased potential of a P2 residue for interactions with the enzyme lowers the rate of deacylation. In contrast to results obtained with chymotrypsin and pancreatic elastase, trypsin is well able to tolerate a P3 proline residue. In the hydrolysis of peptide esters, tissue kallikrein is definitely superior to trypsin. Conversely, peptide bonds are hydrolyzed less efficiently by tissue kallikrein and the acylation reaction is rate-limiting. The influence of the length of peptide substrates is similar in both enzymes and indicates an extension of the substrate recognition site from subsite S3 to at least S'3 of tissue kallikrein and the importance of a hydrogen bond between the P3 carbonyl group and Gly-216 of the enzymes. Tissue kallikrein also tolerates a P3 proline residue well. In sharp contrast to the behaviour of trypsin is the very strong influence of the P2 residue in tissue-kallikrein-catalyzed reactions. kcat/Km varies 75-fold in the series of the dipeptide esters with nonpolar L-amino acid residues in P2, a P2 glycine residue furnishing the worst and phenylalanine the best substrate, whereas this exchange in the pentapeptides changes kcat/Km as much as 730-fold. This behaviour, together with the high value of kcat/Km for Ac-Phe-Arg-OMe of 3.75 X 10(7) M-1 s-1, suggests rate-limiting binding (k1) in the hydrolysis of the best ester substrates.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

6.
Cysteine proteinases are relevant to several aspects of the parasite life cycle and of parasite-host relationship. Moreover, they appear as promising targets for antiparasite chemotherapy. Here, the first quantitative investigation on the steady-state and pre-steady-state kinetics of the papain-like cysteine proteinases from epimastigotes of Trypanosoma cruzi (cruzipain), the agent of Chagas' disease, and from promastigotes of Leishmania infantum, an agent of visceral and cutaneous leishmaniases, is reported. The results indicate that kinetics for the parasite proteinase catalyzed hydrolysis of N-alpha-benzyloxycarbonyl-L-phenylalanyl-L-arginine-(7-amino-4-methylcoumarin) may be consistently fitted to the minimum three-step mechanism involving the acyl.enzyme intermediate E.P: [mechanism: see text] At neutral pH, the k(+3) step (deacylation process) is rate limiting in enzyme catalysis, whereas, at pH<6, the k(+2) step (acylation process) becomes rate limiting. This illustrates the potential danger in interpreting both kcat versus pH profile, given that the acylation or the deacylation step is rate limiting throughout the whole pH range explored, and Km as the true affinity constant for the E:S complex formation. Comparison with the steady-state and pre-steady-state kinetics of homologous plant enzymes suggests that the parasite cysteine proteinase catalytic behavior appears to be of general significance.  相似文献   

7.
The kinetics of nucleotide incorporation into 24/36-mer primer/template DNA by purified fetal calf thymus DNA polymerase (pol) delta was examined using steady-state and pre-steady-state kinetics. The role of the pol delta accessory protein, proliferating cell nuclear antigen (PCNA), on DNA replication by pol delta was also examined by kinetic analysis. The steady-state parameter k(cat) was similar for pol delta in the presence and absence of PCNA (0.36 and 0.30 min(-1), respectively); however, the K(m) for dNTP was 20-fold higher in the absence of PCNA (0.067 versus 1.2 microm), decreasing the efficiency of nucleotide insertion. Pre-steady-state bursts of nucleotide incorporation were observed for pol delta in the presence and absence of PCNA (rates of polymerization (k(pol)) of 1260 and 400 min(-1), respectively). The reduction in polymerization rate in the absence of PCNA was also accompanied by a 2-fold decrease in burst amplitude. The steady-state exonuclease rate of pol delta was 0.56 min(-1) (no burst, 10(3)-fold lower than the rate of polymerization). The small phosphorothioate effect of 2 for correct nucleotide incorporation into DNA by pol delta.PCNA indicated that the rate-limiting step in the polymerization cycle occurs prior to phosphodiester bond formation. A K(d)(dNTP) value of 0.93 microm for poldelta.dNTP binding was determined by pre-steady-state kinetics. A 5-fold increase in K(d)(DNA) for the pol delta.DNA complex was measured in the absence of PCNA. We conclude that the major replicative mammalian polymerase, pol delta, exhibits kinetic behavior generally similar to that observed for several prokaryotic model polymerases, particularly a rate-limiting step following product formation in the steady state (dissociation of oligonucleotides) and a rate-limiting step (probably conformational change) preceding phosphodiester bond formation. PCNA appears to affect pol delta replication in this model mainly by decreasing the dissociation of the polymerase from the DNA.  相似文献   

8.
Haloalkane dehalogenase from Rhodococcus rhodochrous NCIMB 13064 (DhaA) catalyzes the hydrolysis of carbon-halogen bonds in a wide range of haloalkanes. We examined the steady-state and pre-steady-state kinetics of halopropane conversion by DhaA to illuminate mechanistic details of the dehalogenation pathway. Steady-state kinetic analysis of DhaA with a range of halopropanes showed that bromopropanes had higher k(cat) and lower K(M) values than the chlorinated analogues. The kinetic mechanism of dehalogenation was further studied using rapid-quench-flow analysis of 1,3-dibromopropane conversion. This provided a direct measurement of the chemical steps in the reaction mechanism, i.e., cleavage of the carbon-halogen bond and hydrolysis of the covalent alkyl-enzyme intermediate. The results lead to a minimal mechanism consisting of four main steps. The occurrence of a pre-steady-state burst, both for bromide and 3-bromo-1-propanol, suggests that product release is rate-limiting under steady-state conditions. Combining pre-steady-state burst and single-turnover experiments indicated that the rate of carbon-bromine bond cleavage was indeed more than 100-fold higher than the steady-state k(cat). Product release occurred with a rate constant of 3.9 s(-1), a value close to the experimental k(cat) of 2.7 s(-1). Comparing the kinetic mechanism of DhaA with that of the corresponding enzyme from Xanthobacter autotrophicus GJ10 (DhlA) shows that the overall mechanisms are similar. However, whereas in DhlA the rate of halide release represents the slowest step in the catalytic cycle, our results suggest that in DhaA the release of 3-bromo-1-propanol is the slowest step during 1,3-dibromopropane conversion.  相似文献   

9.
Hays H  Berdis AJ 《Biochemistry》2002,41(15):4771-4778
The effect of metal ion substitution on the dynamics of translesion DNA synthesis catalyzed by the bacteriophage T4 DNA polymerase was quantitatively evaluated through steady-state and transient kinetic techniques. Substitution of Mn(2+) for Mg(2+) enhances the steady-state rate of dNMP misinsertion opposite an abasic site by 11-34-fold. At the molecular level, the enhancement in translesion DNA synthesis reflects a substantial increase in the rate of the conformational change preceding phosphoryl transfer for all dNTPs that were tested. This is best illustrated by the biphasic pre-steady-state time course of dAMP insertion opposite an abasic site which indicates that a step after chemistry is rate-limiting for steady-state enzyme turnover. Furthermore, the k(pol) value of 40 s(-1) measured under single-turnover reaction conditions is 20-fold greater than the k(cat) value of 2 s(-1) measured for steady-state enzyme turnover. Finally, the low elemental effect ( approximately 2.4-fold reduction in k(pol)) measured by substituting the alpha-thiotriphosphate analogue for dATP further argues that chemistry is not rate-limiting. In contrast to the biphasic insertion of dAMP, pre-steady-state time courses for the insertion of dCMP, dGMP, or dTMP opposite an abasic site were linear. Nearly identical k(pol) values ( approximately 1 s(-1)) were measured for the insertion of dCMP, dGMP, and dTMP opposite the abasic site using single-turnover conditions. However, the large elemental effects of 27 and 70 measured by substituting the alpha-thiotriphosphate analogues for dCTP and dGTP, respectively, suggest that phosphoryl transfer may be the rate-limiting step for their insertion opposite the abasic site. Various models are discussed in an attempt to explain the effect of metal ion substitution on the dynamics of translesion DNA replication.  相似文献   

10.
Gram-negative bacteria harboring KPC-2, a class A β-lactamase, are resistant to all β-lactam antibiotics and pose a major public health threat. Arg-164 is a conserved residue in all class A β-lactamases and is located in the solvent-exposed Ω-loop of KPC-2. To probe the role of this amino acid in KPC-2, we performed site-saturation mutagenesis. When compared with wild type, 11 of 19 variants at position Arg-164 in KPC-2 conferred increased resistance to the oxyimino-cephalosporin, ceftazidime (minimum inhibitory concentration; 32→128 mg/liter) when expressed in Escherichia coli. Using the R164S variant of KPC-2 as a representative β-lactamase for more detailed analysis, we observed only a modest 25% increase in k(cat)/K(m) for ceftazidime (0.015→0.019 μm(-1) s(-1)). Employing pre-steady-state kinetics and mass spectrometry, we determined that acylation is rate-limiting for ceftazidime hydrolysis by KPC-2, whereas deacylation is rate-limiting in the R164S variant, leading to accumulation of acyl-enzyme at steady-state. CD spectroscopy revealed that a conformational change occurred in the turnover of ceftazidime by KPC-2, but not the R164S variant, providing evidence for a different form of the enzyme at steady state. Molecular models constructed to explain these findings suggest that ceftazidime adopts a unique conformation, despite preservation of Ω-loop structure. We propose that the R164S substitution in KPC-2 enhances ceftazidime resistance by proceeding through "covalent trapping" of the substrate by a deacylation impaired enzyme with a lower K(m). Future antibiotic design must consider the distinctive behavior of the Ω-loop of KPC-2.  相似文献   

11.
The rate-limiting step for hydrolysis of the positively charged oxoester benzoylcholine (BzCh) by human butyrylcholinesterase (BuChE) is deacylation (k(3)), whereas it is acylation (k(2)) for hydrolysis of the homologous thioester benzoylthiocholine (BzSCh). Steady-state hydrolysis of BzCh and BzSCh by wild-type BuChE and its peripheral anionic site mutant D70G was investigated at different hydrostatic pressures, which allowed determination of volume changes associated with substrate binding, and the activation volumes for the chemical steps. A differential nonlinear pressure-dependence of the catalytic parameters for hydrolysis of both substrates by both enzymes was shown. Nonlinearity of the plots may be explained in terms of compressibility changes or rate-limiting changes. To distinguish between these two possibilities, enzyme phosphorylation by diisopropylfluorophosphate (DFP) in the presence of substrate (BzSCh) under pressure was studied. There was no pressure dependence of volume changes for DFP binding or for phosphorylation of either wild-type or D70G. Analysis of the pressure dependence for steady-state hydrolysis of substrates, and for phosphorylation by DFP provided evidence that no enzyme compressibility changes occurred during the catalyzed reactions. Thus, the nonlinear pressure dependence of substrate hydrolysis reflects changes in the rate-limiting step with pressure. Change in rate-determining step occurred at a pressure of 100 MPa for hydrolysis of BzCh by wild-type and at 75 MPa for D70G. For hydrolysis of BzSCh the change occurred at higher pressures because k(2) < k(3) at atmospheric pressure for this substrate. Elementary volume change contributions upon initial binding, productive binding, acylation and deacylation were calculated from the pressure differentiation of kinetic constants. This analysis shed light on the molecular events taking place along the hydrolysis pathways of BzCh and BzSCh by wild-type BuChE and the D70G mutant. In addition, volume change differences between wild-type and D70G provided new evidence that residue D70 in the peripheral site controls hydration of the active site gorge and the dynamics of the water molecule network during catalysis. Finally, a steady-state kinetic study of the oxyanion hole mutant (G117H) showed that substitution of the ethereal sulfur for oxygen in the substrate alters the final adjustment of substrate in the active site and stabilization of the acylation transition state.  相似文献   

12.
Steady-state ATP hydrolysis in the F1-ATPase of the F(O)F1 ATP synthase complex involves rotation of the central gamma subunit relative to the catalytic sites in the alpha3beta3 pseudo-hexamer. To understand the relationship between the catalytic mechanism and gamma subunit rotation, the pre-steady-state kinetics of Mg x ATP hydrolysis in the soluble F1-ATPase upon rapid filling of all three catalytic sites was determined. The experimentally accessible partial reactions leading up to the rate-limiting step and continuing through to the steady-state mode were obtained for the first time. The burst kinetics and steady-state hydrolysis for a range of Mg x ATP concentrations provide adequate constraints for a unique minimal kinetic model that can fit all the data and satisfy extensive sensitivity tests. Significantly, the fits show that the ratio of the rates of ATP hydrolysis and synthesis is close to unity even in the steady-state mode of hydrolysis. Furthermore, the rate of Pi binding in the absence of the membranous F(O) sector is insignificant; thus, productive Pi binding does not occur without the influence of a proton motive force. In addition to the minimal steps of ATP binding, reversible ATP hydrolysis/synthesis, and the release of product Pi and ADP, one additional rate-limiting step is required to fit the burst kinetics. On the basis of the testing of all possible minimal kinetic models, this step must follow hydrolysis and precede Pi release in order to explain burst kinetics. Consistent with the single molecule analysis of Yasuda et al. (Yasuda, R., Noji, H., Yoshida, M., Kinosita, K., and Itoh, H. (2001) Nature 410, 898-904), we propose that the rate-limiting step involves a partial rotation of the gamma subunit; hence, we name this step k(gamma). Moreover, the only model that is consistent with our data and many other observations in the literature suggests that reversible hydrolysis/synthesis can only occur in the active site of the beta(TP) conformer (Abrahams, J. P., Leslie, A. G. W., Lutter, R., and Walker, J. E. (1994) Nature 370, 621-628).  相似文献   

13.
Li Y  Gong Y  Shi G  Blaszczyk J  Ji X  Yan H 《Biochemistry》2002,41(27):8777-8783
6-Hydroxymethyl-7,8-dihydropterin pyrophosphokinase (HPPK) catalyzes the transfer of pyrophosphate from ATP to 6-hydroxymethyl-7,8-dihydropterin (HMDP). Because HPPK is essential for microorganisms but is absent from human and animals, the enzyme is an excellent target for developing antimicrobial agent. Thermodynamic analysis shows that Mg(2+) is important not only for the binding of nucleotides but also for the binding of HMDP. Transient kinetic analysis shows that a step or steps after the chemical transformation are rate-limiting in the reaction catalyzed by HPPK. The pre-steady-state kinetics is composed of a burst phase and a steady-state phase. The rate constant for the burst phase is approximately 50 times larger than that for the steady-state phase. The latter is very similar to the k(cat) value measured by steady-state kinetics. A set of rate constants for the individual steps of the HPPK-catalyzed reaction has been determined by a combination of stopped-flow and quench-flow analyses. These results form a thermodynamic and kinetic framework for dissecting the roles of active site residues in the substrate binding and catalysis by HPPK.  相似文献   

14.
1. The steady-state kinetics of the interconversion of CO2 and HCO3 catalyzed by human carbonic anhydrase C was studied using 1H2O and 2H2O as solvents. The pH-independent parts of the parameters k(cat) and Km are 3-4 times larger in 1H2O than in 2H2O for both directions of the reaction, while the ratios k(cat)/Km show much smaller isotope effects. With either CO2 or HCO3 as substrate the major pH dependence is observed in k(cat), while Km appears independent of pH. The pKa value characterizing the pH-rate profiles is approximately 0.5 unit larger in 2H2O than in 1H2O. 2. The hydrolysis of p-nitrophenyl acetate catalyzed by human carbonic anhudrase C is approximately 35% faster in 2H2O than in 1H2O. In both solvents the pKa values of the pH-rate profiles are similar to those observed for the CO2-HCO3 interconversion. 3. It is tentatively proposed that the rate-limiting step at saturating concentrations of CO2 or HCO3 is an intramolecular proton transfer between two ionizing groups in the active site. It cannot be decided whether the transformation between enzyme-bound CO2 and HCO3 involves a proton trnasfer or not.  相似文献   

15.
The rate-limiting step in the actomyosin adenosinetriphosphatase cycle   总被引:3,自引:0,他引:3  
We have previously shown that myosin does not have to detach from actin during each cycle of ATP hydrolysis. In the present study, using the A-1 isoenzyme of myosin subfragment 1, we have investigated the nature of the rate-limiting steps in the ATPase cycle. Our results show that, at 15 degrees C, at very low ionic strength, KATPase determined from the double-reciprocal plot of ATPase activity vs. actin concentration is more than 6-fold stronger than KBINDING determined by directly measuring the binding of A-1 myosin subfragment 1 to actin during steady-state ATP hydrolysis. Computer modeling shows that this large difference between KATPase and KBINDING is not compatible with Pi release being the rate-limiting step in the ATPase cycle. If Pi release is not rate limiting, it is possible that the ATP hydrolysis step, itself, is rate limiting. However, this predicts that, at high actin concentration, the value of the initial Pi burst should be close to zero. Therefore, we measured the magnitude of the initial Pi burst in the presence of actin, using both direct measurement and measurement of relative fluorescence magnitude. Our results suggest that the magnitude of the initial Pi burst in the presence of actin is considerably higher than would be expected if the ATP hydrolysis step were the rate-limiting step in the ATPase cycle.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

16.
Steady-state kinetics for the hydrolysis of benzoylcholine (BzCh) and benzoylthiocholine (BzSCh) by wild-type human butyrylcholinesterase (BuChE) and by the peripheral anionic site mutant D70G were compared. kcat/Km for the hydrolysis of BzSCh was 17-fold and 32-fold lower than that for hydrolysis of BzCh by wild-type and D70G, respectively. The rate-limiting step for hydrolysis of BzCh was deacylation, whereas acylation was rate-limiting for hydrolysis of BzSCh. Wild-type enzyme and the D70G mutant were found to reach steady-state velocity slowly with BzCh as the substrate. At pH 6, the approach to steady-state for both enzymes consisted of a mono-exponential acceleration upon which a set of damped oscillations was superimposed. From pH 7 to 8.5, the approach to steady-state consisted of a simple exponential acceleration. The damped oscillations were analyzed by both a numerical approximation and simulation based on a theoretical model. BuChE-catalyzed hydrolysis of the thiocholine analogue of BzCh showed neither lags nor oscillations, under the same conditions. The frequency and amplitude of the damped oscillations decreased as the BzCh concentration increased. The apparent induction time for the exponential portion of the lag was calculated from the envelope of the damped oscillations or from the smooth lag. Wild-type BuChE showed a hyperbolic increase in induction time as the BzCh concentration increased (tau max = 210 s at pH 6.0). However, the induction time for D70G was constant over the whole range of BzCh concentrations (tau max = 60 s at pH 6.0). Thus, the induction time does not conform to a simple hysteretic model in which there is a slow conformational transition of the enzyme from an inactive form E to an active form E'. No pH-dependence of the induction time was found between pH 6.0 and 8.5 in sodium phosphate buffers of various concentrations (from 1 mm to 1 m). However, increasing the pH tended to abolish the oscillations (increase the damping factor). This effect was more pronounced for D70G than for wild-type. Although the lyotropic properties of phosphate change from chaotropic at pH 6.0 to kosmotropic at pH > 8.0, no effect of phosphate concentration on the oscillations was noticed at the different pH values, suggesting that the oscillations are not related to a pH-dependent Hofmeister effect of phosphate ions. Simulation and theoretical analysis of the oscillatory behaviour of the approach to the steady-state for BuChE led us to propose a model for the hysteresis of BuChE with BzCh. In this model, the substrate-free enzyme is present as an equilibrium mixture of two forms, E and E'. Substrate binds to E and E', but only Epsilon'S makes products. It is proposed that oscillations originate from a time-dependent change in the local concentration, solvation and/or conformation of substrate in the bulk solution. 1H-NMR measurements provided evidence for a slow equilibrium between two BzCh conformers. Binding of the conformationally preferred substrate conformer leads to products.  相似文献   

17.
Nonphosphorylating nicotinamide adenine dinucleotide (phosphate)- [NAD(P)-] dependent aldehyde dehydrogenases share a number of conserved amino acid residues, several of which are directly implicated in catalysis. In the present study, the role of Glu-268 from nonphosphorylating glyceraldehyde-3-phosphate dehydrogenase (GAPN) from Streptococcus mutans was investigated. Its substitution by Ala resulted in a k(cat) decrease by 3 orders of magnitude. Pre-steady-state analysis showed that, for both the wild-type and E268A GAPNs, the rate-limiting step of the reaction is associated with deacylation. The pH dependence of the rate of acylation of wild-type GAPN is characterized by the contributions of distinct enzyme protonic species with two pK(a)s of 6.2 and 7.5. Substitution of Glu-268 by Ala resulted in a monosigmoidal pH dependence of the rate constant of acylation with a pK(a) of 6.2, which suggested the assignment of pK(a) 7.5 to Glu-268. Moreover, the E268A substitution did not significantly affect the efficiency of acylation of GAPN, showing that Glu-268 is not critically involved in the acylation, which includes Cys-302 nucleophilic activation and hydride transfer. On the contrary, the drastic decrease of the steady-state rate constant for the E268A GAPN demonstrated the essential role of Glu-268 in the deacylation. At basic pH, the solvent isotope effect of 2.3, characterized by a unique pK(a) of 7.7, and the linearity of the proton inventory showed that the rate-limiting process for deacylation is associated with the hydrolysis step and suggested that the glutamate form of Glu-268 acts as a base catalyst in this process. Surprisingly, the double-sigmoidal form of the pH-steady-state rate constant profile, characterized by pK(a) values of 6.1 and 7.4, revealed the high efficiency of the deacylation even at pH lower than 7.4. Therefore, we propose that the major role of Glu-268 is to promote deacylation through activation and orientation of the attacking water molecule, and in addition to act as a base catalyst at basic pH. From these results in relation to those recently described [Marchal, S., and Branlant, G. (1999) Biochemistry 38, 12950-12958], a scenario for the chemical catalysis of GAPN is proposed.  相似文献   

18.
Xia Z  Azurmendi HF  Mildvan AS 《Biochemistry》2005,44(46):15334-15344
The MutT pyrophosphohydrolase, in the presence of Mg2+, catalyzes the hydrolysis of nucleoside triphosphates by nucleophilic substitution at Pbeta, to yield the nucleotide and PP(i). The best substrate for MutT is the mutagenic 8-oxo-dGTP, on the basis of its Km being 540-fold lower than that of dGTP. Product inhibition studies have led to a proposed uni-bi-iso kinetic mechanism, in which PP(i) dissociates first from the enzyme-product complex (k3), followed by NMP (k4), leaving a product-binding form of the enzyme (F) which converts to the substrate-binding form (E) in a partially rate-limiting step (k5) [Saraswat, V., et al. (2002) Biochemistry 41, 15566-15577]. Single- and multiple-turnover kinetic studies of the hydrolysis of dGTP and 8-oxo-dGTP and global fitting of the data to this mechanism have yielded all of the nine rate constants. Consistent with an "iso" mechanism, single-turnover studies with dGTP and 8-oxo-dGTP hydrolysis showed slow apparent second-order rate constants for substrate binding similar to their kcat/Km values, but well below the diffusion limit (approximately 10(9) M(-1) s(-1)): k(on)app = 7.2 x 10(4) M(-1) s(-1) for dGTP and k(on)app = 2.8 x 10(7) M(-1) s(-1) for 8-oxo-dGTP. These low k(on)app values are fitted by assuming a slow iso step (k5 = 12.1 s(-1)) followed by fast rate constants for substrate binding: k1 = 1.9 x 10(6) M(-1) s(-1) for dGTP and k1 = 0.75 x 10(9) M(-1) s(-1) for 8-oxo-dGTP (the latter near the diffusion limit). With dGTP as the substrate, replacing Mg2+ with Mn2+ does not change k1, consistent with the formation of a second-sphere MutT-M2+-(H2O)-dGTP complex, but slows the iso step (k5) 5.8-fold, and its reverse (k(-5)) 25-fold, suggesting that the iso step involves a change in metal coordination, likely the dissociation of Glu-53 from the enzyme-bound metal so that it can function as the general base. Multiple-turnover studies with dGTP and 8-oxo-dGTP show bursts of product formation, indicating partially rate-limiting steps following the chemical step (k2). With dGTP, the slow steps are the chemical step (k2 = 10.7 s(-1)) and the iso step (k5 = 12.1 s(-1)). With 8-oxo-dGTP, the slow steps are the release of the 8-oxo-dGMP product (k4 = 3.9 s(-1)) and the iso step (k5 = 12.1 s(-1)), while the chemical step is fast (k2 = 32.3 s(-1)). The transient kinetic studies are generally consistent with the steady state kcat and Km values. Comparison of rate constants and free energy diagrams indicate that 8-oxo-dGTP, at low concentrations, is a better substrate than dGTP because it binds to MutT 395-fold faster, dissociates 46-fold slower, and has a 3.0-fold faster chemical step. The true dissociation constants (KD) of the substrates from the E-form of MutT, which can now be obtained from k(-1)/k1, are 3.5 nM for 8-oxo-dGTP and 62 microM for dGTP, indicating that 8-oxo-dGTP binds 1.8 x 10(4)-fold tighter than dGTP, corresponding to a 5.8 kcal/mol lower free energy of binding.  相似文献   

19.
The multifunctional protein CAD catalyzes the first three steps in pyrimidine biosynthesis in mammalian cells, including the synthesis of carbamyl phosphate from bicarbonate, MgATP and glutamine. The Syrian hamster CAD glutaminase (GLNase) domain, a trpG-type amidotransferase, catalyzes glutamine hydrolysis in the absence of MgATP and bicarbonate (Km = 95 microM and kcat = 0.14 s-1). Unlike E. coli carbamyl phosphate synthetase (Wellner, V.P., Anderson, P.M., and Meister, A. (1973) Biochemistry 12, 2061-2066), a stable thioester intermediate did not accumulate when the mammalian enzyme was incubated with glutamine. However, a covalent adduct could be isolated when the protein was denatured in acid. The steady state concentration of the intermediate increased with increasing glutamine concentration to nearly one mole per mole of enzyme with half saturation at 105 microM, close to the Km value for glutamine. The adduct formed at the active site of the glutaminase domain. The rate of breakdown of the intermediate (k4), determined directly, was 0.17 s-1 and the rate of formation (k3) was estimated as 0.52 s-1. In the absence of MgATP and bicarbonate, k4 = kcat indicating that the decomposition of the intermediate is the rate-limiting step. The intermediate was chemically and kinetically competent, and the glutamine dissociation constant (330 microM) and rate constants were consistent with steady state kinetics and accurately predicted the steady state concentration of the intermediate. These studies suggest a mechanism similar to the cysteine proteases such as recently proposed by Mei and Zalkin (Mei, B., and Zalkin, H. (1989) J. Biol. Chem. 264, 16613-16619) who identified a catalytic triad in glutamine phosphoribosyl-5'-pyrophosphate amidotransferase, a purF-type enzyme. MgATP and bicarbonate increased kcat of the glutaminase reaction 14-fold by accelerating both the rate of formation and the rate of breakdown of the intermediate, and prevented the accumulation of the intermediate; however, the Km value for glutamine was not significantly altered. The instability of the thioester intermediate leads to appreciable hydrolysis of glutamine in the absence of the other substrates. However, bicarbonate alone spares glutamine by increasing the Km and Ks of glutamine to 600 and 8960 microM, respectively, thus reducing kcat/Km 3-fold when MgATP is limiting. In the absence of MgATP and bicarbonate, ammonia decreased the rate of hydrolysis and the accumulation of the thioester intermediate indicating that ammonia had direct access to the thioester at the GLNase domain active site.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

20.
Dipeptidyl peptidase-IV (DPP-IV) is a serine protease with a signature Asp-His-Ser motif at the active site. Our pH data suggest that Gly-Pro-pNA cleavage catalyzed by DPP-IV is facilitated by an ionization of a residue with a pK of 7.2 +/- 0.1. By analogy to other serine proteases this pK is suggestive of His-Asp assisted Ser addition to the P1 carbonyl carbon of the substrate to form a tetrahedral intermediate. Solvent kinetic isotope effect studies yielded a D2Okcat/Km=2.9+/-0.2 and a D2Okcat=1.7+/-0.2 suggesting that kinetically significant proton transfers contribute to rate limitation during acyl intermediate formation (leaving group release) and hydrolysis. A "burst" of product release during pre steady-state Gly-Pro-pNA cleavage indicated rate limitation in the deacylation half-reaction. Nevertheless, the amplitude of the burst exceeded the enzyme concentration significantly (approximately 15-fold), which is consistent with a branching deacylation step. All of these data allowed us to better understand DPP-IV inhibition by saxagliptin (BMS-477118). We propose a two-step inhibition mechanism wherein an initial encounter complex is followed by covalent intermediate formation. Final inhibitory complex assembly (kon) depends upon the ionization of an enzyme residue with a pK of 6.2 +/- 0.1, and we assigned it to the catalytic His-Asp pair which enhances Ser nucleophilicity for covalent addition. An ionization with a pK of 7.9 +/- 0.2 likely reflects the P2 terminal amine of the inhibitor hydrogen bonding to Glu205/Glu206 in the enzyme active site. The formation of the covalent enzyme-inhibitor complex was reversible and dissociated with a koff of (5.5 +/- 0.4) x 10(-5) s(-1), thus yielding a Ki* (as koff/kon) of 0.35 nM, which is in good agreement with the value of 0.6 nM obtained from steady-state inhibition studies. Proton NMR spectra of DPP-IV showed a downfield resonance at 16.1 ppm. Two additional peaks in the 1H NMR spectra at 17.4 and 14.1 ppm were observed upon mixing the enzyme with saxagliptin. Fractionation factors (phi) of 0.6 and 0.5 for the 17.4 and 14.1 ppm peaks, respectively, are suggestive of short strong hydrogen bonds in the enzyme-inhibitor complex.  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