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 218 毫秒
1.
R L Stein  D A Trainor 《Biochemistry》1986,25(19):5414-5419
The mechanism of inactivation of human leukocyte elastase (HLE) by the chloromethyl ketone MeOSuc-Ala-Ala-Pro-Val-CH2Cl was investigated. The dependence of the first-order rate constant for inactivation on concentration of chloromethyl ketone is hyperbolic and suggests formation of a reversible "Michaelis complex" prior to covalent interaction between the enzyme and inhibitor. However, the observed Ki value is 10 microM, at least 10-fold lower than dissociation constants for complexes formed from interaction of HLE with structurally related substrates or reversible inhibitors, and suggests that Ki is a complex kinetic constant, reflecting the formation and accumulation of both the Michaelis complex and a second complex. It is proposed that this second complex is a hemiketal formed from attack of the active site serine on the carbonyl carbon of the inhibitor. The accumulation of this intermediate may be a general feature of reactions of serine proteases and chloromethyl ketones derived from specific peptides and accounts for the very low Ki values observed for these reactions. The solvent deuterium isotope effect (SIE) on the inactivation step (ki) is 1.58 +/- 0.07 and is consistent with rate-limiting, general-catalyzed attack of the active site His on the methylene carbon of the inhibitor with displacement of chloride anion. The general catalyst is thought to be the active site Asp. In contrast, the SIE on the second-order rate constant for HLE inactivation, ki/Ki, is inverse and equals 0.64 +/- 0.05.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
The crystal structure of the acyl enzyme formed upon inhibition of porcine pancreatic elastase (PPE) by 4-chloro-3-ethoxy-7-guanidinoisocoumarin has been determined at a 1.85-A effective resolution. The chlorine atom is still present in this acyl enzyme, in contrast to the previously reported structure of the 7-amino-4-chloro-3-methoxyisocoumarin-PPE complex where the chlorine atom has been replaced by an acetoxy group. The guanidino group forms hydrogen bonds with the carbonyl group and side-chain hydroxyl group of Thr-41, and the acyl carbonyl group has been twisted out of the oxyanion hole. Molecular modeling indicates that the orientation of the initial Michaelis enzyme-inhibitor complex is quite different from that of the acyl enzyme since simple reconstruction of the isocoumarin ring would result in unfavorable interactions with Ser-195 and His-57. Molecular models were used to design a series of new 7-(alkylureido)- and 7-(alkylthioureido)-substituted derivatives of 3-alkoxy-7-amino-4-chloroisocoumarin as PPE inhibitors. All the 3-ethoxyisocoumarins were better inhibitors than those in the 3-methoxy series due to better interactions with the S1 pocket of PPE. The best ureido inhibitor also contained a tert-butylureido group at the 7-position of the isocoumarin. Due to a predicted interaction with a small hydrophobic pocket on the surface of PPE, this isocoumarin and a related phenylthioureido derivative are among the best irreversible inhibitors thus far reported for PPE (kobs/[I] = 8100 M-1 s-1 and 12,000 M-1 s-1). Kinetic studies of the stability of enzyme-inhibitor complexes suggest that many isocoumarins are alkylating the active site histidine at pH 7.5 via a quinone imine methide intermediate, while at pH 5.0, the predominant pathway appears to be simple formation of a stable acyl enzyme derivative.  相似文献   

3.
Pennati A  Zanetti G  Aliverti A  Gadda G 《Biochemistry》2008,47(11):3418-3425
Despite a number of studies, the formation of the Michaelis complexes between ferredoxin-NADP (+) reductases and NADP(H) eluded detailed investigations by rapid kinetic techniques because of their high formation rates. Moreover, the reversible nature of the reaction of hydride ion transfer between these enzymes and NADPH prevented the obtainment of reliable estimates of the rate constant of the hydride transfer step. Here we show that by working at a high salt concentration, the mechanism of the reaction with NADPH of FprA, a Mycobacterium tuberculosis homologue of adrenodoxin reductase, is greatly simplified, making it amenable to investigation by rapid reaction techniques. The approach presented herein allowed for the first time the observation of the formation of the Michaelis complex between an adrenodoxin reductase-like enzyme and NADPH, and the determination of the related rate constants for association and dissociation. Furthermore, the rate constant for the reaction of hydride ion transfer between NADPH and FAD could be unambiguously assessed. It is proposed that the approach described should be applicable to other ferredoxin reductase enzymes, providing a valuable experimental tool for the study of their kinetic properties.  相似文献   

4.
Zhao G  Jorns MS 《Biochemistry》2006,45(19):5985-5992
Monomeric sarcosine oxidase is a flavoenzyme that catalyzes the oxidation of the methyl group in sarcosine (N-methylglycine). Rapid reaction kinetic studies under anaerobic conditions at pH 8.0 show that the enzyme forms a charge transfer Michaelis complex with sarcosine (E-FAD(ox).sarcosine) that exhibits an intense long-wavelength absorption band (lambda(max) = 516 nm, epsilon(516) = 4800 M(-)(1) cm(-)(1)). Since charge transfer interaction with sarcosine as donor is possible only with the anionic form of the amino acid, the results indicate that the pK(a) of enzyme-bound sarcosine must be considerably lower than the free amino acid (pK(a) = 10.0). No redox intermediate is detectable during sarcosine oxidation, as judged by the isosbestic spectral course observed for conversion of E-FAD(ox).sarcosine to reduced enzyme at 25 or 5 degrees C. The limiting rate of the reductive half-reaction at 25 degrees C (140 +/- 3 s(-)(1)) is slightly faster than turnover (117 +/- 3 s(-)(1)). The kinetics of formation of the Michaelis charge transfer complex can be directly monitored at 5 degrees C where the reduction rate is 4.5-fold slower and complex stability is increased 2-fold. The observed rate of complex formation exhibits a hyperbolic dependence on sarcosine concentration with a finite Y-intercept, consistent with a mechanism involving formation of an initial complex followed by isomerization to yield a more stable complex. Similar results are obtained for charge transfer complex formation with methylthioacetate. The observed kinetics are consistent with structural studies which show that a conformational change occurs upon binding of methylthioacetate and other competitive inhibitors.  相似文献   

5.
The cephalosporin derivatives L 658758 [1-[[3-(acetoxymethyl)-7 alpha-methoxy-8-oxo-5-thia-1-azabicyclo [4.2.0]oct-2-en-2-yl]carbonyl]proline S,S-dioxide] and L 659286 [1-[[7 alpha-methoxy-8-oxo-3-[[(1,2,5,6-tetrahydro-2-methyl-5,6-dioxo- 1,2,4-triazin-3-yl)thio]methyl]-5-thia-1-aza-(6R)-bicyclo[4.2.0]-o ct-2-en-2-yl]carbonyl]pyrrolidine S,S-dioxide] are mechanism based inhibitors of human leukocyte elastase (HLE). The mechanism involves initial formation of a Michaelis complex followed by acylation of the active site serine. The group on the 3'-methylene is liberated during the course of these reactions, followed by partitioning of an intermediate between hydrolysis to regenerate active enzyme and further modification to produce a stable HLE-inhibitor complex. The partition ratio of 2.0 obtained for the reaction with L 658758 approaches that of an optimal inhibitor. These compounds are functionally irreversible inhibitors as the recovery of activity after inactivation is slow. The half-lives at 37 degrees C of the L 658758 and L 659286 derived HLE-I complexes were 9 and 6.5 h, respectively. The complexes produced by both inhibitors are similar chemically since the thermodynamic parameters for activation to regenerate active enzyme are essentially identical. The free energy of activation for this process is dominated primarily by the enthalpy term. The stability of the final complexes likely arises from Michael addition on the active site histidine to the 3'-methylene.  相似文献   

6.
The biosynthesis of uroporphyrinogen III, the precursor of hemes, chlorophylls, corrins and related structures, is catalyzed by the porphobilinogenase system (PBGase), a complex of two enzymes, PBG-Deaminase (PBG-D) and Isomerase. Although the separate enzymes have been studied in some detail less work has been performed on the properties of the complex.In this study the kinetic behaviour of the enzyme PBGase in a normal yeast strain, D273-10B, and its derivative B231 has been investigated. Uroporphyrinogen formation was linear with time up to 2 hr at 37°C. The enzyme complex shows classical Michaelis-Menten kinetics. From the double reciprocal plots kinetic parameters were estimated for PBGase and PBG-D.Porphyrins were found to be competitive inhibitors with respect to porphobilinogen (PBG) and these compounds appeared to act as inhibitors by forming dead-end complexes with the free enzyme. 5-Aminolevulinic acid (ALA) also inhibited PBGase and this inhibition was overcome by addition of levulinic acid (2μM). These results indicate that ALA, is not an inhibitor but acts through its conversion into porphyrins which are the true inhibitors.  相似文献   

7.
Substrate inhibition is a common phenomenon in enzyme chemistry, which is observed only with a fast-reacting substrate enantiomer. We report here for the first time substrate inhibition of an enantioselective enzyme by both substrate enantiomers. The enantioselective substrate inhibition, i.e., different mode of inhibition by each substrate enantiomer, of (S)-specific omega-transaminase was found with various chiral amines. A kinetic model based on ping-pong bi-bi mechanism has been developed and kinetic parameters were measured. The kinetic model reveals that the inhibition by (R)-amine results from formation of Michaelis complex with enzyme-pyridoxal 5'-phosphate, whereas the inhibition by (S)-amine results from the formation of the complex with enzyme-pyridoxamine 5'-phosphate. Substrate inhibition constants (K(SI)) of each (S)-enantiomer of four chiral amines showed a linear correlation with those of cognate (R)-amines. Such a correlation was also found between the K(SI) values and Michaelis constants of (S)-amines. These correlations indicate that recognition mechanisms and active site structures of both enzyme-pyridoxal 5'-phosphate, enzyme-pyridoxamine 5'-phosphate are similar. Taken together with the results, high propensity for non-productive substrate binding strongly suggests that binding pockets of the omega-transaminase is loosely defined, which accounts for the enantioselective substrate inhibition.  相似文献   

8.
Although originally discovered as inhibitors of pencillin-binding proteins, beta-lactams have more recently found utility as serine protease inhibitors. Indeed through their ability to react irreversibly with nucleophilic serine residues they have proved extraordinarily successful as enzyme inhibitors. Consequently there has been much speculation as to the reason for the general effectiveness of beta-lactams as antibacterials or inhibitors of hydrolytic enzymes. The interaction of analogous beta- and gamma-lactams with a serine protease was investigated. Three series of gamma-lactams based upon monocyclic beta-lactam inhibitors of elastase [Firestone, R. A. et al. (1990) Tetrahedron 46, 2255-2262.] but with an extra methylene group inserted between three of the bonds in the ring were synthesized. Their interaction with porcine pancreatic elastase and their efficacy as inhibitors were evaluated through the use of kinetic, NMR, mass spectrometric, and X-ray crystallographic analyses. The first series, with the methylene group inserted between C-3 and C-4 of the beta-lactam template, were readily hydrolyzed but were inactive or very weakly active as inhibitors. The second series, with the methylene group between C-4 and the nitrogen of the beta-lactam template, were inhibitory and reacted reversibly with PPE to form acyl-enzyme complexes, which were stable with respect to hydrolysis. The third series, with the methylene group inserted between C-2 and C-3, were not hydrolyzed and were not inhibitors consistent with lack of binding to PPE. Comparison of the crystal structure of the acyl-enzyme complex formed between PPE and a second series gamma-lactam and that formed between PPE and a peptide [Wilmouth, R. C., et al. (1997) Nat. Struct. Biol. 4, 456-462.] reveals why the complexes formed with this series were resistant to hydrolysis and suggests ways in which stable acyl-enzyme complexes might be obtained from monocyclic gamma-lactam-based inhibitors.  相似文献   

9.
Human beta-casomorphin-7 (NH2-Tyr-Pro-Phe-Val-Glu-Pro-Ile-CO2H) is a naturally occurring peptide inhibitor of elastase that has been shown to form an acyl-enzyme complex stable enough for X-ray crystallographic analysis at pH 5. To investigate the importance of the N-terminal residues of the beta-casomorphin-7 peptide for the inhibition of elastase, kinetic and crystallographic analyses were undertaken to identify the minimum number of residues required for effective formation of a stable complex between truncated beta-casomorphin-7 peptides and porcine pancreatic elastase (PPE). The results clearly demonstrate that significant inhibition of PPE can be effected by simple tri-, tetra-and pentapeptides terminating in a carboxylic acid. These results also suggest that in vivo regulation of protease activity could be mediated via short peptides as well as by proteins. Crystallographic analysis of the complex formed between N-acetyl-Val-Glu-Pro-Ile-CO2H and PPE at pH 5 (to 1.67 A resolution) revealed an active site water molecule in an analogous position to that observed in the PPE/beta-casomorphin-7 structure supportive of its assignment as the 'hydrolytic water' in the deacylation step of serine protease catalysis.  相似文献   

10.
The kinetic mechanism of the oxidative decarboxylation of 2R,3S-isopropylmalate by the NAD-dependent isopropylmalate dehydrogenase of Thermus thermophilus was investigated. Initial rate results typical of random or steady-state ordered sequential mechanisms are obtained for both the wild-type and two mutant enzymes (E87G and E87Q) regardless of whether natural or alternative substrates (2R-malate, 2R,3S-tartrate and/or NADP) are utilized. Initial rate data fail to converge on a rapid equilibrium-ordered pattern despite marked reductions in specificity (kcat/Km) caused by the mutations and alternative substrates. Although the inhibition studies alone might suggest an ordered kinetic mechanism with cofactor binding first, a detailed analysis reveals that the expected noncompetitive patterns appear uncompetitive because the dissociation constants from the ternary complexes are far smaller than those from the binary complexes. Equilibrium fluorescence studies both confirm the random binding of substrates and the kinetic estimates of the dissociation constants of the substrates from the binary complexes. The latter are not distributed markedly by the mutations at site 87. Mutations at site 87 do not affect the dissociation constants from the binary complexes, but do greatly increase the Michaelis constants, indicating that E87 helps stabilize the Michaelis complex of the wild-type enzyme. The available structural data, the patterns of the kinetics results, and the structure of a pseudo-Michaelis complex of the homologous isocitrate dehydrogenase of Escherichia coli suggest that E87 interacts with the nicotinamide ring.  相似文献   

11.
To study the catalytic mechanism of phosphorylation catalyzed by cAMP-dependent protein kinase (PKA) a structure of the enzyme-substrate complex representing the Michaelis complex is of specific interest as it can shed light on the structure of the transition state. However, all previous crystal structures of the Michaelis complex mimics of the PKA catalytic subunit (PKAc) were obtained with either peptide inhibitors or ATP analogs. Here we utilized Ca2+ ions and sulfur in place of the nucleophilic oxygen in a 20-residue pseudo-substrate peptide (CP20) and ATP to produce a close mimic of the Michaelis complex. In the ternary reactant complex, the thiol group of Cys-21 of the peptide is facing Asp-166 and the sulfur atom is positioned for an in-line phosphoryl transfer. Replacement of Ca2+ cations with Mg2+ ions resulted in a complex with trapped products of ATP hydrolysis: phosphate ion and ADP. The present structural results in combination with the previously reported structures of the transition state mimic and phosphorylated product complexes complete the snapshots of the phosphoryl transfer reaction by PKAc, providing us with the most thorough picture of the catalytic mechanism to date.  相似文献   

12.
Serpins (serine protease inhibitors) inhibit target proteases by forming a stable covalent complex in which the cleaved reactive site loop of the serpin is inserted into beta-sheet A of the serpin with concomitant translocation of the protease to the opposite of the initial binding site. Despite recent determination of the crystal structures of a Michaelis protease-serpin complex as well as a stable covalent complex, details on the kinetic mechanism remain unsolved mainly due to difficulties in measuring kinetic parameters of acylation, protease translocation, and deacylation steps. To address the problem, we applied a mathematical model developed on the basis of a suicide inhibition mechanism to the stopped-flow kinetics of fluorescence resonance energy transfer during complex formation between alpha(1)-antitrypsin, a prototype serpin, and proteases. Compared with the hydrolysis of a peptide substrate, acylation of the protease by alpha(1)-antitrypsin is facilitated, whereas deacylation of the acyl intermediate is strongly suppressed during the protease translocation. The results from nucleophile susceptibility of the acyl intermediate suggest strongly that the active site of the protease is already perturbed during translocation.  相似文献   

13.
Bacterial single-stranded (ss) DNA-binding proteins (SSBs) bind and protect ssDNA intermediates formed during cellular DNA replication, recombination and repair reactions. SSBs also form complexes with an array of genome maintenance enzymes via their conserved C-terminal tail (SSB-Ct) elements. In many cases, complex formation with SSB stimulates the biochemical activities of its protein partners. Here, we investigate the mechanism by which Escherichia coli SSB stimulates hydrolysis of ssDNA by Exonuclease I (ExoI). Steady-state kinetic experiments show that SSB stimulates ExoI activity through effects on both apparent k(cat) and K(m). SSB variant proteins with altered SSB-Ct sequences either stimulate more modestly or inhibit ExoI hydrolysis of ssDNA due to increases in the apparent Michaelis constant, highlighting a role for protein complex formation in ExoI substrate binding. Consistent with a model in which SSB stabilizes ExoI substrate binding and melts secondary structures that could impede ExoI processivity, the specific activity of a fusion protein in which ExoI is tethered to SSB is nearly equivalent to that of SSB-stimulated ExoI. Taken together, these studies delineate stimulatory roles for SSB in which protein interactions and ssDNA binding are both important for maximal activity of its protein partners.  相似文献   

14.
Heparin activates the serpin, antithrombin, to inhibit its target blood-clotting proteases by generating new protease interaction exosites. To resolve the effects of these exosites on the initial Michaelis docking step and the subsequent acylation and conformational change steps of antithrombin-protease reactions, we compared the reactions of catalytically inactive S195A and active proteases with site-specific fluorophore-labeled antithrombins that allow monitoring of these reaction steps. Heparin bound to N,N'-dimethyl-N-(acetyl)-N'-(7-nitrobenz-3-oxa-1,3-diazol-4-yl)ethylenediamine (NBD)-fluorophore-labeled antithrombins and accelerated the reactions of the labeled inhibitor with thrombin and factor Xa similar to wild type. Equilibrium binding of NBD-labeled antithrombins to S195A proteases showed that exosites generated by conformationally activating antithrombin with a heparin pentasaccharide enhanced the affinity of the serpin for S195A factor Xa minimally 100-fold. Moreover, additional bridging exosites provided by a hexadecasaccharide heparin activator enhanced antithrombin affinity for both S195A factor Xa and thrombin at least 1000-fold. Rapid kinetic studies showed that these exosite-mediated enhancements in Michaelis complex affinity resulted from increases in k(on) and decreases in k(off) and caused antithrombin-protease reactions to become diffusion-controlled. Competitive binding and kinetic studies with exosite mutant antithrombins showed that Tyr-253 was a critical mediator of exosite interactions with S195A factor Xa; that Glu-255, Glu-237, and Arg-399 made more modest contributions to these interactions; and that exosite interactions reduced k(off) for the Michaelis complex interaction. Together these results show that exosites generated by heparin activation of antithrombin function both to promote the formation of an initial antithrombin-protease Michaelis complex and to favor the subsequent acylation of this complex.  相似文献   

15.
4-Nitrophenyl-N-substituted carbamates (1-6) are the pseudo-substrate inhibitors of porcine pancreatic cholesterol esterase. Thus, the first step of the inhibition (Ki step) is the formation of the enzyme inhibitor tetrahedral adduct and the second step of the inhibition (kc) is the formation of the carbamyl enzyme. The formation of the enzyme inhibitor tetrahedral adduct is further divided into two steps, the formation of the enzyme-inhibitor complex with the dissociation constant, KS, at the first step and the formation of the enzyme-inhibitor tetrahedral adduct from the complex at the second step. The two-step mechanism for the formation of the enzyme-inhibitor tetrahedral adduct is confirmed by the pre-steady-state kinetics. The results of quantitative structure-activity relationships for the pre-steady-state inhibitions of cholesterol esterase by carbamates 1-6 indicate that values of -logKs and logk2/k-2 are correlated with the Taft substituent constant, sigma*, and the rho* values from these correlations are -0.33 and 0.1, respectively. The negative rho* value for the -logKS-sigma*-correlation indicates that the first step of the two-step formation of the enzyme-inhibitor tetrahedral adduct (KS step) is the formation of the positive enzyme inhibitor complex. The positive rho* value for the logk2/k-2 -sigma*-correlation indicates that the enzyme inhibitor tetrahedral adduct is more negative than the enzyme inhibitor complex. Finally, the two-step mechanism for the formation of the enzyme inhibitor tetrahedral adduct is proposed according to these results. Thus, the partially positive charge is developed at nitrogen of carbamates 1-6 in the enzyme-inhibitor complex probably due to the hydrogen bonding between the lone pair of nitrogen of carbamates 1-6 and the amide hydrogen of the oxyanion hole of the enzyme. The second step of the two-step formation of the enzyme-inhibitor tetrahedral adduct is the nucleophilic attack of the serine of the enzyme to the carbonyl group of carbamates 1-6 in the enzyme-inhibitor complex and develops the negative-charged oxygen in the adduct.  相似文献   

16.
Kovaleva EG  Plapp BV 《Biochemistry》2005,44(38):12797-12808
Binding of NAD+ to wild-type horse liver alcohol dehydrogenase is strongly pH-dependent and is limited by a unimolecular step, which may be related to a conformational change of the enzyme-NAD+ complex. Deprotonation during binding of NAD+ and inhibitors that trap the enzyme-NAD+ complex was examined by transient kinetics with pH indicators, and formation of complexes was monitored by absorbance and protein fluorescence. Reactions with pyrazole and trifluoroethanol had biphasic proton release, whereas reaction with caprate showed proton release followed by proton uptake. Proton release (200-550 s(-1)) is a common step that precedes binding of all inhibitors. At all pH values studied, the rate constants for proton release or uptake matched those for formation of ternary complexes, and the most significant quenching of protein fluorescence (or perturbation of adenine absorbance at 280 nm) was observed for enzyme species involved in deprotonation steps. Kinetic simulations of the combined transient data for the multiple signals indicate that all inhibitors bind faster and tighter to the unprotonated enzyme-NAD+ complex, which has a pK of about 7.3. The results suggest that rate-limiting deprotonation of the enzyme-NAD+ complex is coupled to the conformational change and controls the formation of ternary complexes.  相似文献   

17.
We report here the x-ray structure of a covalent serpin-proteinase complex, alpha1-proteinase inhibitor (alpha1PI) with porcine pancreatic elastase (PPE), which differs from the only other x-ray structure of such a complex, that of alpha1PI with trypsin, in showing nearly complete definition of the proteinase. alpha1PI complexes with trypsin, PPE, and human neutrophil elastase (HNE) showed similar rates of deacylation and enhanced susceptibility to proteolysis by exogenous proteinases in solution. The differences between the two x-ray structures therefore cannot arise from intrinsic differences in the inhibition mechanism. However, self-proteolysis of purified complex resulted in rapid cleavage of the trypsin complex, slower cleavage of the PPE complex, and only minimal cleavage of the HNE complex. This suggests that the earlier alpha1 PI-trypsin complex may have been proteolyzed and that the present structure is more likely to be representative of serpin-proteinase complexes. The present structure shows that active site distortion alone is sufficient for inhibition and suggests that enhanced proteolysis is not necessarily exploited in vivo.  相似文献   

18.
Recent studies on taurine:alpha-ketoglutarate dioxygenase (TauD) from Escherichia coli have provided evidence for a three-step, minimal kinetic mechanism involving the quaternary TauD.Fe(II).alpha-ketoglutarate.taurine complex, the taurine-hydroxylating Fe(IV)-oxo intermediate (J) that forms upon reaction of the quaternary complex with O(2), and a poorly defined, Fe(II)-containing intermediate state that converts in the rate-limiting step back to the quaternary complex [Price, J. C., Barr, E. W., Tirupati, B., Bollinger, J. M., Jr., and Krebs, C. (2003) Biochemistry 42, 7497-7508]. The mapping of this kinetic mechanism onto the consensus chemical mechanism for the Fe(II)- and alpha-ketoglutarate-dependent engendered several predictions and additional questions that have been experimentally addressed in the present study. The results demonstrate (1) that postulated intermediates between the quaternary complex and J accumulate very little or not at all; (2) that decarboxylation of alpha-ketoglutarate occurs prior to or concomitantly with formation of J; (3) that the second intermediate state comprises one or more product complex with Mossbauer features that are partially resolved from those of the binary TauD.Fe(II), ternary TauD.Fe(II).alpha-ketoglutarate, and quaternary TauD.Fe(II).alpha-ketoglutarate.taurine complexes; and (4) that the rate-determining step in the catalytic cycle is release of product(s) prior to the rapid, ordered binding of alpha-ketoglutarate and then taurine to regenerate the O(2)-reactive quaternary complex. The results thus integrate the previously proposed kinetic and chemical mechanisms and indicate which of the postulated intermediates in the latter will be detectable only upon perturbation of the kinetics by changes in reaction conditions (e.g., temperature), protein mutagenesis, the use of substrate analogues, or some combination of these.  相似文献   

19.
Liu W  Rogers CJ  Fisher AJ  Toney MD 《Biochemistry》2002,41(41):12320-12328
The kinetics of inhibition of dialkylglycine decarboxylase by five aminophosphonate inhibitors are presented. Two of these [(R)-1-amino-1-methylpropanephosphonate and (S)-1-aminoethanephosphonate] are slow binding inhibitors. The inhibitors follow a mechanism in which a weak complex is rapidly formed, followed by slow isomerization to the tight complex. Here, the tight complexes are bound 10-fold more tightly than the weak, initial complexes. The slow onset inhibition occurs with t(1/2) values of 1.3 and 0.55 min at saturating inhibitor concentrations for the AMPP and S-AEP inhibitors, respectively, while dissociation of these inhibitor complexes occurs with t(1/2) values of 13 and 4.6 min, respectively. The X-ray structures of four of the inhibitors in complex with dialkylglycine decarboxylase have been determined to resolutions ranging from 2.6 to 2.0 A, and refined to R-factors of 14.5-19.5%. These structures show variation in the active site structure with inhibitor side chain size and slow binding character. It is proposed that the slow binding behavior originates in an isomerization from an initial complex in which the PLP pyridine nitrogen-D243 OD2 distance is approximately 2.9 A to one in which it is approximately 2.7 A. The angles that the C-P bonds make with the p orbitals of the aldimine pi system are correlated with the reactivities of the analogous amino acid substrates, suggesting a role for stereoelectronic effects in Schiff base reactivity.  相似文献   

20.
Kinetic studies on the pig heart cytoplasmic malate dehydrogenase have been performed over a wide range of conditions using the full time course of the reaction and computer simulation to obtain the kinetic parameters. The maximum velocity and Michaelis constants for the oxidation of reduced coenzyme have been determined as a fundtion of pH in 0.05 M phosphate buffer at 15 degrees. At pH 7.5 and at low substrate concentrations, the kinetic data are consistent with a sequential addition of substrates, coenzyme binding first, and involving the formation of at least one ternary complex. No oxalacetate binding to the enzyme was observed. The rate constants for the dissociation of coenzyme from the enzyme-coenzyme complex are small enough to define the maximum velocity in either direction of the reaction. These data, plus data using deuterated reduced coenzyme, indicate that the chemical transformation step is not rate determining. It is also shown that DPNH binding can be tight enough to practically exclude the possibility of obtaining initial velocities when measuring the reduction of DPN. Kinetic abnormalities do appear at higher substrate or product concentrations, but these do not appear to be related to the formation of inactive abortice, complexes.  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