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 31 毫秒
1.
Mycobacterium tuberculosis alpha-isopropylmalate synthase (MtIPMS) catalyzes the condensation of acetyl-coenzyme A (AcCoA) with alpha-ketoisovalerate (alpha-KIV) and the subsequent hydrolysis of alpha-isopropylmalyl-CoA to generate the products CoA and alpha-isopropylmalate (alpha-IPM). This is the first committed step in l-leucine biosynthesis. We have purified recombinant MtIPMS and characterized it using a combination of steady-state kinetics, isotope effects, isotopic labeling, and (1)H-NMR spectroscopy. The alpha-keto acid specificity of the enzyme is narrow, and the acyl-CoA specificity is absolute for AcCoA. In the absence of alpha-KIV, MtIPMS does not enolize the alpha protons of AcCoA but slowly hydrolyzes acyl-CoA analogues. Initial velocity studies, product inhibition, and dead-end inhibition studies indicate that MtIPMS follows a nonrapid equilibrium random bi-bi kinetic mechanism, with a preferred pathway to the ternary complex. MtIPMS requires two catalytic bases for maximal activity (both with pK(a) values of ca. 6.7), and we suggest that one catalyzes deprotonation and enolization of AcCoA and the other activates the water molecule involved in the hydrolysis of alpha-isopropylmalyl-CoA. Primary deuterium and solvent kinetic isotope effects indicate that there is a step after chemistry that is rate-limiting, although, with poor substrates such as pyruvate, hydrolysis becomes partially rate-limiting. Our data is inconsistent with the suggestion that a metal-bound water is involved in hydrolysis. Finally, our results indicate that the hydrolysis of alpha-isopropylmalyl-CoA is direct, without the formation of a cyclic anhydride intermediate. On the basis of these results, a chemical mechanism for the MtIPMS-catalyzed reaction is proposed.  相似文献   

2.
3.
Catalytic mechanism of hamster arylamine N-acetyltransferase 2   总被引:1,自引:0,他引:1  
Wang H  Liu L  Hanna PE  Wagner CR 《Biochemistry》2005,44(33):11295-11306
Arylamine N-acetyltransferases (NATs) catalyze an acetyl group transfer from AcCoA to primary arylamines, hydrazines, and hydrazides and play a very important role in the metabolism and bioactivation of drugs, carcinogens, and other xenobiotics. The reaction follows a ping-pong bi-bi mechanism. Structure analysis of bacterial NATs revealed a Cys-His-Asp catalytic triad that is strictly conserved in all known NATs. Previously, we have demonstrated by kinetic and isotope effect studies that acetylation of the hamster NAT2 is dependent on a thiolate-imidazolium ion pair (Cys-S(-)-His-ImH(+)) and not a general acid-base catalysis. In addition, we established that, after formation of the acetylated enzyme intermediate, the active-site imidazole, His-107, is likely deprotonated at physiological pH. In this paper, we report steady-state kinetic studies of NAT2 with two acetyl donors, acetyl coenzyme A (AcCoA) and p-nitrophenyl acetate (PNPA), and four arylamine substrates. The pH dependence of k(cat)/K(AcCoA) exhibited two inflection points at 5.32 +/- 0.13 and 8.48 +/- 0.24, respectively. The pK(a) at 5.32 is virtually identical with the previously reported pK(a) of 5.2 for enzyme acetylation, reaffirming that the first half of the reaction is catalyzed by a thiolate-imidazolium ion pair in the active site. The inflection point at 8.48 indicates that a pH-sensitive group on NAT2 is involved in AcCoA binding. A Br?nsted plot constructed by the correlation of log k(4) and log k(H)2(O) with the pK(a) for each arylamine substrate and water displays a linear free-energy relationship in the pK(a) range from -1.7 (H(2)O) to 4.67 (PABA), with a slope of beta(nuc) = 0.80 +/- 0.1. However, a further increase of the pK(a) from 4.67 (PABA) to 5.32 (anisidine) resulted in a 2.5-fold decrease in the k(4) value. Analysis of the pH-k(cat)/K(PABA) profile revealed a pK(a) of 5.52 +/- 0.14 and a solvent kinetic isotope effect (SKIE) of 2.01 +/- 0.04 on k(cat)/K(PABA). Normal solvent isotope effects of 4.8 +/- 0.1, 3.1 +/- 0.1, and 3.2 +/- 0.1 on the k(cat)/K(b) for anisidine, pABglu, and PNA, respectively, were also determined. These observations are consistent with a deacetylation mechanism dominated by nucleophilic attack of the thiol ester for arylamines with pK(a) values or=5.5. The general base is likely His-107 because the His-107 to Gln and Asn mutants were found to be devoid of catalytic activity. In contrast, an increase in pH-dependent hydrolysis of the acetylated enzyme was not observed over a pH range of 5.2-7.5. On the basis of these observations, a catalytic mechanism for the acetylation of arylamines by NAT2 is proposed.  相似文献   

4.
The homocitrate synthase from Thermus thermophilus (TtHCS) is a metal-activated enzyme with either Mg(2+) or Mn(2+) capable of serving as the divalent cation. The enzyme exhibits a sequential kinetic mechanism. The mechanism is steady state ordered with α-ketoglutarate (α-Kg) binding prior to acetyl-CoA (AcCoA) with Mn(2+), whereas it is steady state random with Mg(2+), suggesting a difference in the competence of the E·Mn·α-Kg·AcCoA and E·Mg·α-Kg·AcCoA complexes. The mechanism is supported by product and dead-end inhibition studies. The primary isotope effect obtained with deuterioacetylCoA (AcCoA-d(3)) in the presence of Mg(2+) is unity (value 1.0) at low concentrations of AcCoA, whereas it is 2 at high concentrations of AcCoA. Data suggest the presence of a slow conformational change induced by binding of AcCoA that accompanies deprotonation of the methyl group of AcCoA. The solvent kinetic deuterium isotope effect is also unity at low AcCoA, but is 1.7 at high AcCoA, consistent with the proposed slow conformational change. The maximum rate is pH independent with either Mg(2+) or Mn(2+) as the divalent metal ion, whereas V/K(α-Kg) (with Mn(2+)) decreases at low and high pH giving pK values of about 6.5 and 8.0. Lysine is a competitive inhibitor that binds to the active site of TtHCS, and shares some of the same binding determinants as α-Kg. Lysine binding exhibits negative cooperativity, indicating cross-talk between the two monomers of the TtHCS dimer. Data are discussed in terms of the overall mechanism of TtHCS.  相似文献   

5.
Malate synthase, an enzyme of the glyoxylate pathway, catalyzes the condensation and subsequent hydrolysis of acetyl-coenzyme A (acetyl-CoA) and glyoxylate to form malate and CoA. In the present study, we present the 1.95 A-resolution crystal structure of Escherichia coli malate synthase isoform G in complex with magnesium, pyruvate, and acetyl-CoA, and we compare it with previously determined structures of substrate and product complexes. The results reveal how the enzyme recognizes and activates the substrate acetyl-CoA, as well as conformational changes associated with substrate binding, which may be important for catalysis. On the basis of these results and mutagenesis of active site residues, Asp 631 and Arg 338 are proposed to act in concert to form the enolate anion of acetyl-CoA in the rate-limiting step. The highly conserved Cys 617, which is immediately adjacent to the presumed catalytic base Asp 631, appears to be oxidized to cysteine-sulfenic acid. This can explain earlier observations of the susceptibility of the enzyme to inactivation and aggregation upon X-ray irradiation and indicates that cysteine oxidation may play a role in redox regulation of malate synthase activity in vivo. There is mounting evidence that enzymes of the glyoxylate pathway are virulence factors in several pathogenic organisms, notably Mycobacterium tuberculosis and Candida albicans. The results described in this study add insight into the mechanism of catalysis and may be useful for the design of inhibitory compounds as possible antimicrobial agents.  相似文献   

6.
The steady-state kinetic parameters for epimerization of UDP-galactose by UDP-galactose 4-epimerase from Escherichia coli (GalE), Y149F-GalE, and S124A-GalE have been measured as a function of pH. The deuterium kinetic isotope effects for epimerization of UDP-galactose-C-d(7) by these enzymes have also been measured. The results show that the activity of wild-type GalE is pH-independent in the pH range of 5.5-9.3, and there is no significant deuterium kinetic isotope effect in the reaction of UDP-galactose-C-d(7). It is concluded that the rate-limiting step for epimerization by wild-type GalE is not hydride transfer and must be either a diffusional process or a conformational change. Epimerization of UDP-galactose-C-d(7) by Y149F-GalE proceeds with a pH-dependent deuterium kinetic isotope effect on k(cat) of 2.2 +/- 0.4 at pH 6.2 and 1.1 +/- 0.5 at pH 8.3. Moreover, the plot of log k(cat)/K(m) breaks downward on the acid side with a fitted value of 7.1 for the pK(a). It is concluded that the break in the pH-rate profile arises from a change in the rate-limiting step from hydride transfer at low pH to a conformational change at high pH. Epimerization of UDP-galactose-C-d(7) by S124A-GalE proceeds with a pH-independent deuterium kinetic isotope effect on k(cat) of 2.0 +/- 0.2 between pH 6 and 9. Both plots of log k(cat) and log k(cat)/K(m) display pH dependence. The plot of log k(cat) versus pH breaks downward with a pK(a) of 6.35 +/- 0.10. The plot of log k(cat)/K(m) versus pH is bell-shaped, with fitted pK(a) values of 6.76 +/- 0.09 and 9.32 +/- 0.21. It is concluded that hydride transfer is rate-limiting, and the pK(a) of 6.7 for free S124A-GalE is assigned to Tyr 149, which displays the same value of pK(a) when measured spectrophotometrically in this variant. Acid-base catalysis by Y149F-GalE is attributed to Ser 124, which is postulated to rescue catalysis of proton transfer in the absence of Tyr 149. The kinetic pK(a) of 7.1 for free Y149F-GalE is lower than that expected for Ser 124, as proven by the pH-dependent kinetic isotope effect. Epimerization by the doubly mutated Y149F/S124A-GalE proceeds at a k(cat) that is lower by a factor of 10(7) than that of wild-type GalE. This low rate is attributed to the synergistic actions of Tyr 149 and Ser 124 in wild-type GalE and to the absence of any internal catalysis of hydride transfer in the doubly mutated enzyme.  相似文献   

7.
Qian J  West AH  Cook PF 《Biochemistry》2006,45(39):12136-12143
Homocitrate synthase (acetyl-coenzyme A:2-ketoglutarate C-transferase; E.C. 2.3.3.14) catalyzes the condensation of AcCoA and alpha-ketoglutarate to give homocitrate and CoA. The enzyme was found to be a Zn-containing metalloenzyme using inductively coupled plasma mass spectrometry. Dead-end analogues of alpha-ketoglutarate were used to obtain information on the topography of the alpha-ketoglutarate binding site. The alpha-carboxylate and alpha-oxo groups of alpha-ketoglutarate are required for optimum binding to coordinate to the active site Zn. Optimum positioning of the alpha-carboxylate, alpha-oxo, and gamma-carboxylate of alpha-ketoglutarate is likely mimicked by the location in space of the 2-carboxylate, pyridine nitrogen, and 4 carboxylate of pyridine 2,4-dicarboxylate. The pH dependence of the kinetic parameters was determined to obtain information on the chemical mechanism of homocitrate synthase. The V profile is bell shaped with slopes of 1 and -1, giving pKa values of 6.7 and 8.0, while V/K(AcCoA) exhibits a slope of 2 on the acidic side with an average pKa value of 6.6 and a slope of -2 on basic side of the profile with an average pKa value of 8.2. The V/K(alpha-Kg) pH-rate profile exhibits a single pKa of 6.9 on the acidic side and two on the basic side with an average value of 7.8. The pH dependence of the Ki for glyoxylate, a competitive inhibitor vs alpha-ketoglutarate, gives a pKa of 7.1 for a group, required to be protonated for optimum binding. Data suggest a chemical mechanism for the enzyme in which alpha-ketoglutarate first binds to the active site Zn via its alpha-carboxylate and alpha-oxo groups, followed by acetyl-CoA. A general base then accepts a proton from the methyl of acetyl-CoA, and a general acid protonates the carbonyl of alpha-ketoglutarate in the formation of homocitryl-CoA. The general acid then acts as a base in deprotonating Zn-OH2 in the hydrolysis of homocitryl-CoA to give homocitrate and CoA. A solvent deuterium kinetic isotope effect of 1 is measured for homocitrate synthase, while a small pH-independent primary kinetic deuterium isotope effect (approximately 1.3) is observed using deuterioacetyl-CoA. Data suggest rate-limiting condensation to form the alkoxide of homocitryl-CoA, followed by hydrolysis to give products.  相似文献   

8.
Lee SS  Yu S  Withers SG 《Biochemistry》2003,42(44):13081-13090
The unusual enzyme, Gracilariopsis alpha-1,4-glucan lyase of the sequence-related glycoside hydrolase family 31, cleaves the glycosidic bond of alpha-1,4-glucans via a beta-elimination reaction involving a covalent glycosyl-enzyme intermediate (Lee, S. S., Yu, S., and Withers, S. G. (2002) J. Am. Chem. Soc. 124, 4948-4949). The classical bell-shaped pH dependence of k(cat)/K(m) indicates two ionizable groups in the active site with apparent pK(a) values of 3.05 and 6.66. Br?nsted relationships of log k(cat) versus pK(a) and log(k(cat)/K(m)) versus pK(a) for a series of aryl glucosides both show a linear monotonic dependence on leaving group pK(a) with low beta(lg) values of 0.32 and 0.33, respectively. The combination of these low beta(lg) values with large secondary deuterium kinetic isotope effects (k(H)/k(D) = 1.16 - 1.19) on the first step indicate a glycosylation step with substantial glycosidic bond cleavage and proton donation to the leaving group oxygen at the transition state. Developed oxocarbenium ion character of the transition state is also suggested by the potent inhibition afforded by acarbose and 1-deoxynojirimycin (K(i) = 20 and 130 nM, respectively) and by the substantial rate reduction afforded by adjacent fluorine substitution. For only one substrate, 5-fluoro-alpha-D-glucopyranosyl fluoride, was the second elimination step shown to be rate-limiting. The large alpha-secondary deuterium kinetic isotope effect (k(H)/k(D) = 1.23) at C-1 and the small primary deuterium kinetic isotope effect (k(H)/k(D) = 1.92) at C-2 confirm an E2 mechanism with strong E1 character for this second step. This considerable structural and mechanistic similarity with retaining alpha-glucosidases is clear evidence for the evolution of an enzyme mechanism within the family.  相似文献   

9.
10.
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   

11.
Although aldolase-catalyzed condensations proceed by stepwise mechanisms via the intermediacy of nucleophilic enol(ate)s or enamines, the mechanisms of those enzymes that catalyze Claisen-type condensations are unclear. The reaction pathway followed by an enzyme from this second group, malate synthase, has been studied by the double-isotope fractionation method to determine whether the reaction is stepwise or concerted. In agreement with earlier work, a deuterium kinetic isotope effect D(V/K) of 1.3 +/- 0.1 has been found when [2H3]acetyl-CoA is the substrate. The 13C isotope effect at the aldehydic carbon of glyoxylate has also been measured. For this determination, the malate product (containing the carbon of interest at C-2) was quantitatively transformed into a new sample of malate having the carbon of interest at C-4. This material was decarboxylated by malic enzyme to produce the appropriate CO2 for isotope ratio mass spectrometric analysis. The 13C isotope effect with [1H3]acetyl-CoA [that is, 13(V/K)H] is 1.0037 +/- 0.0004. By use of the known values of the intermolecular and intramolecular deuterium effects and of 13(V/K)H, the value of the 13C isotope effect when deuteriated [2H3]acetyl-CoA is the substrate [that is, 13(V/K)D] can be predicted for three possible mechanisms. If 13(V/K)H is a kinetic isotope effect and the reaction is concerted, the value of the 13C effect on deuteriation of acetyl-CoA will rise to 1.011; if 13(V/K)H is a kinetic isotope effect and the reaction is stepwise, the value of the 13C effect will fall to 1.0025; and if the 13C effect is an equilibrium isotope effect deriving from glyoxylate dehydration, the reaction is necessarily stepwise, and the value of 13(V/K)D will be 1.0037, unchanged from that of 13(V/K)H. Experimentally, the value of 13(V/K)D is 1.0037 +/- 0.0007, which requires that malate synthase follow a stepwise path. It is therefore clear that the two salient characteristics of enzymes that catalyze Claisen-like condensations, namely, the absence of enzyme-catalyzed proton exchange with solvent and the inversion of the configuration at the nucleophilic center, which had been suggestive of a concerted pathway, are not mechanistically diagnostic.  相似文献   

12.
Qian J  Khandogin J  West AH  Cook PF 《Biochemistry》2008,47(26):6851-6858
Homocitrate synthase (acetyl-coenzyme A: 2-ketoglutarate C-transferase; E.C. 2.3.3.14) (HCS) catalyzes the condensation of acetyl-CoA (AcCoA) and alpha-ketoglutarate (alpha-KG) to give homocitrate and CoA. Although the structure of an HCS has not been solved, the structure of isopropylmalate synthase (IPMS), a homologue, has been solved (Koon, N., Squire, C. J., and Baker, E. N. (2004) Proc. Natl. Acad. Sci. U.S.A. 101, 8295-8300). Three active site residues in IPMS, Glu-218, His-379, and Tyr-410, were proposed as candidates for catalytic residues involved in deprotonation of the methyl group of AcCoA prior to the Claisen condensation to give homocitrylCoA. All three of the active site residues in IPMS are conserved in the HCS from Saccharomyces cerevisiae. Site-directed mutagenesis has been carried out to probe the role of the homologous residues, Glu-155, His-309, and Tyr-320, in the S. cerevisiae HCS. No detectable activity was observed for the H309A and H309N mutant enzyme, but a slight increase in activity was observed for H309A in the presence of 300 mM imidazole, which is still 1000-fold lower than that of wild type (wt). The E155Q and E155A mutant enzymes exhibited 1000-fold lower activity than wt. The activity of E155A, but not of E155Q, could be partially rescued by formate; a K act of 60 mM with a modest 4-fold maximum activation was observed. In the presence of formate, E155A gives k cat, K AcCoA, and K alpha-KG values of 0.0031 s (-1), 13 muM, and 39 microM, respectively, while a primary kinetic deuterium isotope effect of about 1.4 was obtained on V, with deuterium in the methyl of AcCoA. The pH dependence of k cat for E155A in the presence of formate gave a p K a of 7.9 for a group that must be protonated for optimum activity, similar to that observed for the wt enzyme. However, a partial change was observed on the acid side of the profile, compared to the all or none change observed for wt giving a p K a of about 6.7. The k cat for E155Q decreased at high pH, similar to the wt enzyme, but was pH independent at low pH. The Y320F mutant enzyme only lost 25-fold activity compared to that of the wt, giving k cat, K AcCoA, and K alpha-KG values of 0.039 s (-1), 33 microM, and 140 microM, respectively, and a primary kinetic deuterium isotope effect of 1.3 and 1.8 on V/ K AcCoA and V, respectively; the pH dependence of k cat was similar to that of the wt. These data, combined with a constant pH molecular dynamics simulation study, suggest that a catalytic dyad comprising Glu-155 and His-309 acts to deprotonate the methyl group of AcCoA, while Tyr320 is likely not directly involved in catalysis, but may aid in orienting the reactant and/or the catalytic dyad.  相似文献   

13.
Polgár L 《Biochemistry》1999,38(47):15548-15555
Oligopeptidase B, a member of the novel prolyl oligopeptidase family of serine peptidases, is involved in cell invasion by trypanosomes. The kinetic analysis of the reactions of oligopeptidase B, which preferentially cleaves peptides at two adjacent basic residues, has revealed significant differences from the trypsin-like serine peptidases. (i) The pH dependence of k(cat)/K(m) deviates from normal bell-shaped curves due to ionization of an enzymatic group characterized by a macroscopic pK(a) of approximately 8.3. The effect of this group is abolished at high ionic strength. (ii) The second-order acylation rate constants, k(cat)/K(m), are similar with the ester and the corresponding amide substrates, suggesting that their chemical reactivity does not prevail in the rate-limiting step. The kinetic deuterium isotope effects indicate that the rate-limiting step for k(cat)/K(m) is principally governed by conformational changes. (iii) The pH-k(cat)/K(m) profile and the very low rate constant for benzoyl-citrulline ethyl ester reveal a new kinetically influential group ionizing below the pK(a) of the active site histidine and indicate that the positive charge of arginine is essential for effective catalysis. (iv) The enzyme is inhibited by high concentrations of substrate. The mechanism of inhibition markedly varies with the reaction conditions. (v) The optimum temperature for the reactions of amide substrates is unusually low, slightly below 25 degrees C, whereas with benzoyl-arginine ethyl ester a linear Eyring plot is obtained up to 39 degrees C. The positive entropies of activation point to substantial reorganization of water molecules upon substrate binding.  相似文献   

14.
Sikora AL  Frankel BA  Blanchard JS 《Biochemistry》2008,47(40):10781-10789
Arylamine N-acetyltransferases (NATs) are cytosolic enzymes that catalyze the transfer of the acetyl group from acetyl coenzyme A (AcCoA) to the free amino group of arylamines and hydrazines. Previous studies have reported that overexpression of NAT from Mycobacterium smegmatis and Mycobacterium tuberculosis may be responsible for increased resistance to the front-line antitubercular drug, isoniazid, by acetylating and hence inactivating the prodrug. We report the kinetic characterization of M. tuberculosis NAT which reveals that substituted anilines are excellent substrates but that isoniazid is a very poor substrate for this enzyme. We propose that the expression of NAT from M. tuberculosis (TBNAT) is unlikely to be a significant cause of isoniazid resistance. The kinetic parameters for a variety of TBNAT substrates were examined, including 3-amino-4-hydroxybenzoic acid and AcCoA, revealing K m values of 0.32 +/- 0.03 and 0.14 +/- 0.02 mM, respectively. Steady-state kinetic analysis of TBNAT reveals that the enzyme catalyzes the reaction via a bi-bi ping-pong kinetic mechanism. The pH dependence of the kinetic parameters reveals that one enzyme group must be deprotonated for optimal catalytic activity and that two amino acid residues at the active site of the free enzyme are involved in binding and/or catalysis. Solvent kinetic isotope effects suggest that proton transfer steps are not rate-limiting in the overall reaction for substituted aniline substrates but become rate-limiting when poor hydrazide substrates are used.  相似文献   

15.
Flagg SC  Giri N  Pektas S  Maroney MJ  Knapp MJ 《Biochemistry》2012,51(33):6654-6666
Prolyl hydroxylase domain 2 (PHD2) is deemed a primary oxygen sensor in humans, yet many details of its underlying mechanism are still not fully understood. (Fe(2+) + αKG)PHD2 is 6-coordinate, with a 2His/1Asp facial triad occupying three coordination sites, a bidentate α-ketoglutarate occupying two sites, and an aquo ligand in the final site. Turnover is thought to be initiated upon release of the aquo ligand, creating a site for O(2) to bind at the iron. Herein we show that steady-state turnover is faster under acidic conditions, with k(cat) exhibiting a kinetic pK(a) = 7.22. A variety of spectroscopic probes were employed to identify the active-site acid, through comparison of (Fe(2+) + αKG)PHD2 at pH 6.50 with pH 8.50. The near-UV circular dichroism spectrum was virtually unchanged at elevated pH, indicating that the secondary structure did not change as a function of pH. UV-visible and Fe X-ray absorption spectroscopy indicated that the primary coordination sphere of Fe(2+) changed upon increasing the pH; extended X-ray absorption fine structure analysis found a short Fe-(O/N) bond length of 1.96 ? at pH 8.50, strongly suggesting that the aquo ligand was deprotonated at this pH. Solvent isotope effects were measured during steady-sate turnover over a wide pH-range, with an inverse solvent isotope effect (SIE) of k(cat) observed ((D(2)O)k(cat) = 0.91 ± 0.03) for the acid form; a similar SIE was observed for the basic form of the enzyme ((D(2)O)k(cat) = 0.9 ± 0.1), with an acid equilibrium offset of ΔpK(a) = 0.67 ± 0.04. The inverse SIE indicated that aquo release from the active site Fe(2+) immediately precedes a rate-limiting step, suggesting that turnover in this enzyme may be partially limited by the rate of O(2) binding or activation, and suggesting that aquo release is relatively slow. The unusual kinetic pK(a) further suggested that PHD2 might function physiologically to sense both intracellular pO(2) as well as pH, which could provide for feedback between anaerobic metabolism and hypoxia sensing.  相似文献   

16.
Johnson CM  Huang B  Roderick SL  Cook PF 《Biochemistry》2004,43(49):15534-15539
The pH dependence of kinetic parameters was determined in both reaction directions to obtain information about the acid-base chemical mechanism of serine acetyltransferase from Haemophilus influenzae (HiSAT). The maximum rates in both reaction directions, as well as the V/K(serine) and V/K(OAS), decrease at low pH, exhibiting a pK of approximately 7 for a single enzyme residue that must be unprotonated for optimum activity. The pH-independent values of V(1)/E(t), V(1)/K(serine)E(t), V/K(AcCoA)E(t), V(2)/E(t), V(2)/K(OAS)E(t), and V/K(CoA)E(t) are 3300 +/- 180 s(-1), (9.6 +/- 0.4) x 10(5) M(-1) s(-1), 3.3 x 10(6) M(-1) s(-1), 420 +/- 50 s(-1), (2.1 +/- 0.5) x 10(4) M(-1) s(-1), and (4.2 +/- 0.7) x 10(5) M(-1) s(-1), respectively. The K(i) values for the competitive inhibitors glycine and l-cysteine are pH-independent. The solvent deuterium kinetic isotope effects on V and V/K in the direction of serine acetylation are 1.9 +/- 0.2 and 2.5 +/- 0.4, respectively, and the proton inventories are linear for both parameters. Data are consistent with a single proton in flight in the rate-limiting transition state. A general base catalytic mechanism is proposed for the serine acetyltransferase. Once acetyl-CoA and l-serine are bound, an enzymic general base accepts a proton from the l-serine side chain hydroxyl as it undergoes a nucleophilic attack on the carbonyl of acetyl-CoA. The same enzyme residue then functions as a general acid, donating a proton to the sulfur atom of CoASH as the tetrahedral intermediate collapses, generating the products OAS and CoASH. The rate-limiting step in the reaction at limiting l-serine levels is likely formation of the tetrahedral intermediate between serine and acetyl-CoA.  相似文献   

17.
The P450 2E1-catalyzed oxidation of ethanol to acetaldehyde is characterized by a kinetic deuterium isotope effect that increases K(m) with no effect on k(cat), and rate-limiting product release has been proposed to account for the lack of an isotope effect on k(cat) (Bell, L. C., and Guengerich, F. P. (1997) J. Biol. Chem. 272, 29643-29651). Acetaldehyde is also a substrate for P450 2E1 oxidation to acetic acid, and k(cat)/K(m) for this reaction is at least 1 order of magnitude greater than that for ethanol oxidation to acetaldehyde. Acetic acid accounts for 90% of the products generated from ethanol in a 10-min reaction, and the contribution of this second oxidation has been overlooked in many previous studies. The noncompetitive intermolecular kinetic hydrogen isotope effects on acetaldehyde oxidation to acetic acid ((H)(k(cat)/K(m))/(D)(k(cat)/K(m)) = 4.5, and (D)k(cat) = 1.5) are comparable with the isotope effects typically observed for ethanol oxidation to acetaldehyde, and k(cat) is similar for both reactions, suggesting a possible common catalytic mechanism. Rapid quench kinetic experiments indicate that acetic acid is formed rapidly from added acetaldehyde (approximately 450 min(-1)) with burst kinetics. Pulse-chase experiments reveal that, at a subsaturating concentration of ethanol, approximately 90% of the acetaldehyde intermediate is directly converted to acetic acid without dissociation from the enzyme active site. Competition experiments suggest that P450 2E1 binds acetic acid and acetaldehyde with relatively high K(d) values, which preclude simple tight binding as an explanation for rate-limiting product release. The existence of a rate-determining step between product formation and release is postulated. Also proposed is a conformational change in P450 2E1 occurring during the course of oxidation and the discrimination of P450 2E1 between acetaldehyde and its hydrated form, the gem-diol. This multistep P450 reaction is characterized by kinetic control of individual reaction steps and by loose binding of all ligands.  相似文献   

18.
Hydroxysteroid dehydrogenases (HSDs) are essential for the biosynthesis and mechanism of action of all steroid hormones. We report the complete kinetic mechanism of a mammalian HSD using rat 3alpha-HSD of the aldo-keto reductase superfamily (AKR1C9) with the substrate pairs androstane-3,17-dione and NADPH (reduction) and androsterone and NADP(+) (oxidation). Steady-state, transient state kinetics, and kinetic isotope effects reconciled the ordered bi-bi mechanism, which contained 9 enzyme forms and permitted the estimation of 16 kinetic constants. In both reactions, loose association of the NADP(H) was followed by two conformational changes, which increased cofactor affinity by >86-fold. For androstane-3,17-dione reduction, the release of NADP(+) controlled k(cat), whereas the chemical event also contributed to this term. k(cat) was insensitive to [(2)H]NADPH, whereas (D)k(cat)/K(m) and the (D)k(lim) (ratio of the maximum rates of single turnover) were 1.06 and 2.06, respectively. Under multiple turnover conditions partial burst kinetics were observed. For androsterone oxidation, the rate of NADPH release dominated k(cat), whereas the rates of the chemical event and the release of androstane-3,17-dione were 50-fold greater. Under multiple turnover conditions full burst kinetics were observed. Although the internal equilibrium constant favored oxidation, the overall K(eq) favored reduction. The kinetic Haldane and free energy diagram confirmed that K(eq) was governed by ligand binding terms that favored the reduction reactants. Thus, HSDs in the aldo-keto reductase superfamily thermodynamically favor ketosteroid reduction.  相似文献   

19.
Ralph EC  Fitzpatrick PF 《Biochemistry》2005,44(8):3074-3081
N-Methyltryptophan oxidase (MTOX), a flavoenzyme from Escherichia coli, catalyzes the oxidative demethylation of secondary amino acids such as N-methyltryptophan or N-methylglycine (sarcosine). MTOX is one of several flavin-dependent amine oxidases whose chemical mechanism is still debated. The kinetic properties of MTOX with the slow substrate sarcosine were determined. Initial rate data are well-described by the equation for a ping-pong kinetic mechanism, in that the V/K(O)()2 value is independent of the sarcosine concentration at all accessible concentrations of oxygen. The k(cat)/K(sarc) pH profile is bell-shaped, with pK(a) values of 8.8 and about 10; the latter value matches the pK(a) value of the substrate nitrogen. The k(cat) pH profile exhibits a single pK(a) value of 9.1 for a group that must be unprotonated for catalysis. There is no significant solvent isotope effect on the k(cat)/K(sarc) value. With N-methyl-(2)H(3)-glycine as the substrate, there is a pH-independent kinetic isotope effect on k(cat), k(cat)/K(sarc), and the rate constant for flavin reduction, with an average value of 7.2. Stopped-flow spectroscopy with both the protiated and deuterated substrate failed to detect any intermediates between the enzyme-substrate complex and the fully reduced enzyme. These results are used to evaluate proposed chemical mechanisms.  相似文献   

20.
Guinea pig liver transglutaminase (TGase) reacts with 0.1 mM N-Cbz-L-Glu(gamma-p-nitrophenyl ester)Gly (5, prepared herein, K(M) = 0.02 mM) to undergo rapid acylation that can be followed spectrophotometrically at 400 nm (pH 7.0, 25 degrees C). Deacylation of the transiently formed thiolester acyl enzyme intermediate via catalytic aminolysis was studied in the presence of six primary amines of widely varying basicity (pK(NH+) = 5.6-10.5). Steady-state kinetic studies were performed to measure k(cat) and K(M) values for each amine substrate. A Br?nsted plot constructed through the correlation of log(k(cat)/K(M)) and pK(NH+) for each amine substrate displays a linear free-energy relationship with a slope beta(nuc) = -0.37 +/- 0.08. The shallow negative slope is consistent with a general-base-catalyzed deacylation mechanism in which a proton is removed from the amine substrate during its rate-limiting nucleophilic attack on the thiolester carbonyl. Kinetic isotope effects were measured for four acceptor substrates (water, kie = 1.1 +/- 0.1; aminoacetonitrile, kie = 5.9 +/- 1.2; glycine methyl ester, kie = 3.4 +/- 0.7; N-Ac-L-lysine methyl ester, kie = 1.1 +/- 0.1) and are consistent with a proton in flight at the rate-limiting transition state. The active site general-base implicated by these kinetic results is believed to be His-334, of the highly conserved TGase Cys-His-Asp catalytic triad.  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