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.
Daubner SC  Fitzpatrick PF 《Biochemistry》1999,38(14):4448-4454
The active site of tyrosine hydroxylase consists of a hydrophobic cleft with an iron atom near the bottom. Within the cleft are several charged residues which are conserved across the family of pterin-dependent hydroxylases. We have studied four of these residues, glutamates 326 and 332, aspartate 328, and arginine 316 in tyrosine hydroxylase, by site-directed substitution with alternate amino acid residues. Replacement of arginine 316 with lysine results in a protein with a Ktyr value that is at least 400-fold greater and a V/Ktyr value that is 4000-fold lower than those found in the wild-type enzyme; substitution with alanine, serine, or glutamine yields insoluble enzyme. Arginine 316 is therefore critical for the binding of tyrosine. Replacement of glutamate 326 with alanine has no effect on the KM value for tyrosine and results in a 2-fold increase in the KM value for tetrahydropterin. The Vmax for DOPA production is reduced 9-fold, and the Vmax for dihydropterin formation is reduced 4-fold. These data suggest that glutamate 326 is not directly involved in catalysis. Replacement of aspartate 328 with serine results in a 26-fold higher KM value for tyrosine, a 8-fold lower Vmax for dihydropterin formation, and a 13-fold lower Vmax for DOPA formation. These data suggest that aspartate 328 has a role in tyrosine binding. Replacement of glutamate 332 with alanine results in a 10-fold higher KM value for 6-methyltetrahydropterin with no change in the KM value for tyrosine, a 125-fold lower Vmax for DOPA formation, and an only 3.3-fold lower Vmax for tetrahydropterin oxidation. These data suggest that glutamate 332 is required for productive tetrahydropterin binding.  相似文献   

2.
Tyrosine hydroxylase catalyzes the tetrahydropterin-dependent hydroxylation of tyrosine to form 3,4-dihydroxyphenylalanine. Several nonphysiological aromatic amino acids have been examined as inhibitors and substrates for bovine adrenal tyrosine hydroxylase. The Ki values for para-substituted phenylalanines increase as the size of the substituent increases. For each A2 increase in surface area of the substituent, the free energy of binding becomes 50 cal more positive. Replacement of the phenyl ring with a pyridyl ring decreases the affinity about one order of magnitude. A number of these aromatic amino acids are also substrates for the enzyme. The KM values again increase in size with increasing size of the substituent, but the Vmax value is independent of the reactivity of the amino acid. The effect of size on binding is consistent with a tight interaction between the para position region of the substrate and the enzyme. The lack of a change in the Vmax value is consistent with the rate-limiting step in catalysis by bovine tyrosine hydroxylase being formation of the hydroxylating intermediate rather than hydroxylation of the amino acid. These results will be useful in designing mechanism-based inhibitors of catecholamine biosynthesis and establish that the mechanisms of rat and bovine tyrosine hydroxylase do not differ significantly.  相似文献   

3.
The interaction between cytochrome P-450 and NADPH-cytochrome c reductase during catalysis has been investigated with a reconstituted monooxygenase system composed of the two purified enzyme components and synthetic phospholipid. Steady state kinetic data are consistent with a scheme in which the formation of a binary complex between the two proteins precedes catalysis. The formation of this binary complex is described by a simple mass action equation. In agreement with this equation, the observed Vmax for benzphetamine N-demethylation was found to be directly proportional to the calculated concentration of the cytochrome P-450 . reductase complex. Furthermore, with appropriate reductase/cytochrome P-450 mole ratios, the Vmax could be shown to be linearly dependent on either the reductase or the cytochrome P-450 concentration alone. In contrast, the Km parameter is independent of the complex concentration, indicating that no change in the rate-limiting step has occurred. Thus a distinction should be made between a rate-limiting enzyme component and the rate-limiting step in this multienzyme system.  相似文献   

4.
The aza analogue (RS)-3-hydroxy-2,5-pyrrolidinedione-3-acetic acid (6) of the five-membered citric anhydride (2) was prepared in the sequence citric acid----2-phenyl-1,3-dioxolan-4-one-5,5-diacetic acid (1)----citric acid beta-amide (3)----6 and used to resolve ambiguities in the mechanism of the citrate synthase reaction. The results yield no indication for the formation of anhydride 2 on the enzyme and favour the direct hydrolysis of the intermediate (3S)-citryl-CoA. Ammonolysis of the dioxolanone 1 in the reaction sequence described above produced not only citric acid beta-amide but also the alpha-isomer. This is shown to originate in the transient formation of anhydride 2. Hydrolysis of the dioxolanone 1 under "physiological conditions" occurs via anhydride 2, generated in intramolecular bifunctional catalysis by a protonated and a deprotonated carboxyl group. The catalytic residue Asp375 of citrate synthase is considered to operate on the enzyme as does the protonated carboxyl group in the chemical reaction and to generate enolic acetyl-CoA in cooperative catalysis with His274. This reaction of Asp375 may also facilitate the hydrolysis of citryl-CoA.  相似文献   

5.
We examined the effects of hypoxia (8% O2) on in vivo tyrosine hydroxylation, a rate-limiting step for catecholamine synthesis, in the rat adrenal gland. The hydroxylation rate was determined by measuring the rate of accumulation of 3,4-dihydroxyphenylalanine (DOPA) after decarboxylase inhibition. One hour after hypoxic exposure, DOPA accumulation decreased to 60% of control values, but within 2 h it doubled. At 2 h, the apparent Km values for tyrosine and for biopterin cofactor of tyrosine hydroxylase (TH) in the soluble fraction were unchanged, whereas the Vmax value increased by 30%. The content of total or reduced biopterin was unchanged, but the content of tyrosine increased by 80%. Tyrosine administration had little effect on DOPA accumulation under room air conditions but enhanced DOPA accumulation under hypoxia. After denervation of the adrenal gland, the hypoxia-induced increase in DOPA accumulation and in the Vmax value was abolished, whereas the hypoxia-induced increase in tyrosine content was persistent. These results suggest that in vivo tyrosine hydroxylation is enhanced under hypoxia, although availability of oxygen is reduced. The enhancement is the result of both an increase in tyrosine content coupled with increased sensitivity of TH to changes in tyrosine tissue content and of an increase in dependence of TH on tyrosine levels. The increase in the sensitivity of TH and in the Vmax value is neurally induced, whereas the increase in tyrosine content is regulated by a different mechanism.  相似文献   

6.
The hydroxylation rate and rate of tyrosine catabolism are measured by injection of ring-deuterated L-phenylalanine and ring-deuterated L-tyrosine and subsequent deuterium determination in the water fraction of the blood. The hydroxylation rate was confirmed as the rate-limiting step. Whereas tyrosine catabolism yields normal Michaelis-Menten kinetics, homotropic activation is demonstrated for the hydroxylation step. In all tumor rats under study, the rates of phenylalanine hydroxylation and tyrosine catabolism are decreased. The delta-values to control increase with tumor age. In tumor rats, the hydroxylation step remains the rate-limiting one. Kinetic data indicate that the decrease in hydroxylation is due to a decreased turnover rate of the enzyme, whereas the increase in tyrosine catabolism is caused by the branching off of tyrosine or an intermediate on the route to D2O formation. The results are discussed with respect to altered levels of tetrahydrobiopterin, phenylalanine and tyrosine found during parallel investigations.  相似文献   

7.
P M Weiss  C Y Chen  W W Cleland  P F Cook 《Biochemistry》1988,27(13):4814-4822
We have used deuterium and 15N isotope effects to study the relative rates of the steps in the mechanisms of alanine and glutamate dehydrogenases. The proposed chemical mechanisms for these enzymes involve carbinolamine formation, imine formation, and reduction of the imine to the amino acid [Grimshaw, C.E., Cook, P.F., & Cleland, W.W. (1981) Biochemistry 20, 5655; Rife, J.E., & Cleland, W.W. (1980) Biochemistry 19, 2328]. These steps are almost equally rate limiting for V/Kammonia with alanine dehydrogenase, while with glutamate dehydrogenase carbinolamine formation, imine formation, and release of glutamate after hydride transfer provide most of the rate limitation of V/Kammonia. Release of oxidized nucleotide is largely rate limiting for Vmax for both enzymes. When beta-hydroxypyruvate replaces pyruvate, or 3-acetylpyridine NADH (Acpyr-NADH) or thio-NADH replaces NADH with alanine dehydrogenase, nucleotide release no longer limits Vmax, and hydride transfer becomes more rate limiting. With glutamate dehydrogenase, replacement of alpha-ketoglutarate by alpha-ketovalerate makes hydride transfer more rate limiting. Use of Acpyr-NADPH has a minimal effect with alpha-ketoglutarate but causes an 8-fold decrease in Vmax with alpha-ketovalerate, with hydride transfer the major rate-limiting step. In contrast, thio-NADPH with either alpha-keto acid causes carbinolamide formation to become almost completely rate limiting. These studies show the power of multiple isotope effects in deducing details of the chemistry and changes in rate-limiting step(s) in complicated reaction mechanisms such as those of alanine and glutamate dehydrogenases.  相似文献   

8.
4-trans-(N,N-dimethylamino)cinnamaldehyde (DACA) is a chromophoric and fluorogenic substrate of aldehyde dehydrogenase. Fluorescence of DACA is enhanced by binding to aldehyde dehydrogenase in the absence of catalysis both in the presence and absence of the coenzyme analogue 5'AMP. DACA binds to aldehyde dehydrogenase with a dissociation constant of 1-3 microM and stoichiometry of 2 mol mol(-1) enzyme. Incorporation of DACA during catalysis was also investigated and found to be 2 mol DACA mol(-1) enzyme. Effect of pH on the stoichiometry of DACA incorporation during catalysis has shown that DACA incorporation remained constant at 2 mol DACA mol(-1) enzyme, despite a 74-fold velocity enhancement between pH 5.0 and 9.0. Increase of pH increased decomposition of enzyme-acyl intermediate without affecting the rate-limiting step of the reaction. At pH 7.0 the pH stimulated velocity enhancement was 10-fold over that at pH 5.0; further velocity enhancement (11.5-fold that of pH 7.0) was achieved by 150 microM Mg(2+) ions. The velocity at pH 7.0 with Mg(2+) exceeded that of pH 9.0, and that at maximal pH stimulation at pH 9.5. It was observed that level of intermediate decreased to about 1 mol mol(-1) enzyme, indicating that Mg(2+) ions increased the rate of decomposition of the enzyme-acyl intermediate and shifted the rate-limiting step of the reaction to another step in the reaction sequence.  相似文献   

9.
The 4-hydroxybenzoyl-CoA (4-HB-CoA) thioesterase from Pseudomonas sp. strain CBS3 catalyzes the final step of the 4-chlorobenzoate degradation pathway, which is the hydrolysis of 4-HB-CoA to coenzyme A (CoA) and 4-hydroxybenzoate (4-HB). In previous work, X-ray structural analysis of the substrate-bound thioesterase provided evidence of the role of an active site Asp17 in nucleophilic catalysis [Thoden, J. B., Holden, H. M., Zhuang, Z., and Dunaway-Mariano, D. (2002) X-ray crystallographic analyses of inhibitor and substrate complexes of wild-type and mutant 4-hydroxybenzoyl-CoA thioesterase. J. Biol. Chem. 277, 27468-27476]. In the study presented here, kinetic techniques were used to test the catalytic mechanism that was suggested by the X-ray structural data. The time course for the multiple-turnover reaction of 50 μM [(14)C]-4-HB-CoA catalyzed by 10 μM thioesterase supported a two-step pathway in which the second step is rate-limiting. Steady-state product inhibition studies revealed that binding of CoA (K(is) = 250 ± 70 μM; K(ii) = 900 ± 300 μM) and 4-HB (K(is) = 1.2 ± 0.2 mM) is weak, suggesting that product release is not rate-limiting. A substantial D(2)O solvent kinetic isotope effect (3.8) on the steady-state k(cat) value (18 s(-1)) provided evidence that a chemical step involving proton transfer is the rate-limiting step. Taken together, the kinetic results support a two-chemical pathway. The microscopic rate constants governing the formation and consumption of the putative aspartyl 17-(4-hydroxybenzoyl)anhydride intermediate were determined by simulation-based fitting of a kinetic model to time courses for the substrate binding reaction (5.0 μM 4-HB-CoA and 0.54 μM thioesterase), single-turnover reaction (5 μM [(14)C]-4-HB-CoA catalyzed by 50 μM thioesterase), steady-state reaction (5.2 μM 4-HB-CoA catalyzed by 0.003 μM thioesterase), and transient-state multiple-turnover reaction (50 μM [(14)C]-4-HB-CoA catalyzed by 10 μM thioesterase). Together with the results obtained from solvent (18)O labeling experiments, the findings are interpreted as evidence of the formation of an aspartyl 17-(4-hydroxybenzoyl)anhydride intermediate that undergoes rate-limiting hydrolytic cleavage at the hydroxybenzoyl carbonyl carbon atom.  相似文献   

10.
Oxidation of methionine into methionine sulfoxide is associated with many pathologies and is described to exert regulatory effects on protein functions. Two classes of methionine sulfoxide reductases, called MsrA and MsrB, have been described to reduce the S and the R isomers of the sulfoxide of methionine sulfoxide back to methionine, respectively. Although MsrAs and MsrBs display quite different x-ray structures, they share a similar, new catalytic mechanism that proceeds via the sulfenic acid chemistry and that includes at least three chemical steps with 1) the formation of a sulfenic acid intermediate and the concomitant release of methionine; 2) the formation of an intra-disulfide bond; and 3) the reduction of the disulfide bond by thioredoxin. In the present study, it is shown that for the Neisseria meningitidis MsrA, 1) the rate-limiting step is associated with the reduction of the Cys-51/Cys-198 disulfide MsrA bond by thioredoxin; 2) the formation of the sulfenic acid intermediate is very efficient, thus suggesting catalytic assistance via amino acids of the active site; 3) the rate-determining step in the formation of the Cys-51/Cys-198 disulfide bond is that leading to the formation of the sulfenic intermediate on Cys-51; and 4) the apparent affinity constant for methionine sulfoxide in the methionine sulfoxide reductase step is 80-fold higher than the Km value determined under steady-state conditions.  相似文献   

11.
Homogeneous thromboxane synthase from human platelets converted prostaglandin H2 (PGH2) to thromboxane A2 (measured as thromboxane B2, TxB2), 12(L)-hydroxy-5,8,10-heptadecatrienoic acid (HHT), and malondialdehyde (MDA) in equimolar amounts under a variety of experimental conditions. PGG2 was transformed to MDA and corresponding 15- and 12-hydroperoxy products. PGH1 was enzymatically transformed into 12(L)-hydroxy-8,10-heptadecadienoic acid (HHD) and PGH3 into TxB3 and 12(L)-hydroxy-5,8,10,14-heptadecatetraenoic acid (delta 14-HHT) as earlier reported for solubilized and partially purified thromboxane synthase preparations. The ratio of thromboxane to C17 hydroxy fatty acid formation was 1:1 with PGG2, PGH2, and PGH3 as substrates. These results confirm and extend earlier observations with partially purified enzyme that the three products are formed in a common enzymatic pathway (Diczfalusy, U., Falardeau, P., and Hammarstr?m, S. (1977) FEBS Lett. 84, 271-274). A convenient spectrophotometric assay for thromboxane synthase activity measuring the ultraviolet light absorption of the C17 hydroxy acid formed (e.g., HHT) was developed. The validity of the assay was determined employing specific inhibitors for thromboxane synthase. The substrate specificity of thromboxane synthase was determined using this assay. PGG2 and PGH3 showed Vmax and KM values similar to those of PGH2. The KM value of PGH1 was also identical to that of PGH2 but the Vmax value PGH1 was more than twice as high as that of PGH2.  相似文献   

12.
DuBois JL  Klinman JP 《Biochemistry》2005,44(34):11381-11388
The copper amine oxidases (CAOs) catalyze the O(2)-dependent, two-electron oxidation of amines to aldehydes at an active site that contains Cu(II) and topaquinone (TPQ) cofactor. TPQ arises from the autocatalytic, post-translational oxidation of a tyrosine side chain in the active site. Monooxygenation within the ring of tyrosine at a single Cu(II) site is unique in biology and occurs as an early step in the formation of TPQ. The mechanism of this reaction has been further examined in the CAO from Hansenula polymorpha (HPAO). When a Clark electrode fitted to a custom-made, gastight apparatus over a range of initial concentrations of O(2) was used, rates of O(2) consumption at levels greater than air are seen to be reduced relative to earlier results, yielding K(D)(apparent) = 216 microM for O(2). This is consistent with a mechanism in which O(2) binds reversibly to the active site, triggering a conformational change that promotes ligation of tyrosinate to Cu(II). The activated Cu(II)-tyrosinate species has been proposed to react with O(2) in a rate-limiting step, although it was also possible that breakdown of a putative peroxy-intermediate controlled TPQ formation. To test the latter hypothesis, Cu(II)-free HPAO was prepared with 3,5-ring-[(2)H(2)]-tyrosine incorporated throughout the primary sequence. The absence of an isotope effect on the rate of TPQ formation eliminates cleavage of this C-H bond in a proposed Cu(II)-aryl-peroxide intermediate as a rate limiting step. The role of methionine 634, previously found to moderate O(2) binding during the catalytic cycle, is shown here to serve a similar function in TPQ formation. As with catalysis, the rate of TPQ formation correlates with the volume of the hydrophobic side chain at position 634, implicating similar binding sites for O(2) during catalysis and cofactor biogenesis.  相似文献   

13.
Although the activation of low-molecular weight protein tyrosine phosphatases by certain purines and purine derivatives was first described three decades ago, the mechanism of this rate enhancement was unknown. As an example, adenine activates the yeast low-molecular weight protein tyrosine phosphatase LTP1 more than 30-fold. To examine the structural and mechanistic basis of this phenomenon, we have determined the crystal structure of yeast LTP1 complexed with adenine. In the crystal structure, an adenine molecule is found bound in the active site cavity, sandwiched between the side chains of two large hydrophobic residues at the active site. Hydrogen bonding to the side chains of other active site residues, as well as some water-mediated hydrogen bonds, also helps to fix the position of the bound adenine molecule. An ordered water was found in proximity to the bound phosphate ion present in the active site, held by hydrogen bonding to N3 of adenine and Odelta1 of Asp-132. On the basis of the crystal structure, we propose that this water molecule is the nucleophile that participates in the dephosphorylation of the phosphoenzyme intermediate. Solvent isotope effect studies show that there is no rate-determining transfer of a solvent-derived proton in the transition state for the dephosphorylation of the phosphoenzyme intermediate. Such an absence of general base catalysis of water attack is consistent with the stability of the leaving group, namely, the thiolate anion of Cys-13. Consequently, adenine activates the enzyme by binding and orienting a water nucleophile in proximity to the phosphoryl group of the phosphoenzyme intermediate, thus increasing the rate of the dephosphorylation step, a step that is normally the rate-limiting step of this enzymatic reaction.  相似文献   

14.
Radiolabeling of human liver alpha-L-fucosidase (alpha-L-fucoside fucohydrolase, EC 3.2.1.51) with [1-3H]conduritol C trans-epoxide revealed that there are four active sites per tetrameric enzyme complex. Solvent isotope effect experiments give evidence for a proton transfer at the rate-limiting step in catalysis. Transglycosylase activity was observed using methanol as an alternative glycone acceptor to produce methyl alpha-L-fucoside, suggesting that alpha-L-fucose is formed when water is the acceptor. Initial burst kinetics experiments suggest that a glycosyl-enzyme intermediate is formed, although the magnitude of the burst is not stoichiometric with the number of active sites. These data, along with previous results, suggest a general acid-general base catalytic mechanism involving double inversion of stereochemistry at C-1 of fucose, as well as the formation of either a covalent glycosyl-enzyme intermediate or a tight ion pair between a charged active-site residue and a hypothetical fucosyl oxocarbonium ion intermediate.  相似文献   

15.
Mechanism for the peroxynitrite scavenging activity by anthocyanins   总被引:4,自引:0,他引:4  
Tsuda T  Kato Y  Osawa T 《FEBS letters》2000,484(3):207-210
We show that anthocyanins can function as potent inhibitors of the formation of nitrated tyrosine in vitro, and clarified how pelargonidin (Pel), which has a mono-hydroxyl group on the B-ring, can scavenge peroxynitrite (ONOO(-)) by detection of the reaction products. Pel was reacted with ONOO(-), then the reaction mixture was analyzed using high-performance liquid chromatography (HPLC). The HPLC analyses showed two novel peaks assumed to be the reaction products. Based on the instrumental analyses, the reaction products were identified as p-hydroxybenzoic acid and 4-hydroxy-3-nitrobenzoic acid. Pel can protect tyrosine from undergoing nitration through the formation of p-hydroxybenzoic acid and 4-hydroxy-3-nitrobenzoic acid.  相似文献   

16.
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   

17.
JNK3 alpha 1 is predominantly a neuronal specific MAP kinase that is believed to require, like all MAP kinases, both threonine and tyrosine phosphorylation for maximal enzyme activity. In this study we investigated the in vitro activation of JNK3 alpha 1 by MAP kinase kinase 4 (MKK4), MAP kinase kinase 7 (MKK7), and the combination of MKK4 + MKK7. Mass spectral analysis showed that MKK7 was capable of monophosphorylating JNK3 alpha 1 in vitro, whereas both MKK4 and MKK7 were required for bisphosphorylation and maximal enzyme activity. Measuring catalysis under Vmax conditions showed MKK4 + MKK7-activated JNK3 alpha 1 had Vmax 715-fold greater than nonactivated JNK3 alpha 1 and MKK7-activated JNK3 alpha 1 had Vmax 250-fold greater than nonactivated JNK3 alpha 1. In contrast, MKK4-activated JNK3 alpha 1 had no increase in Vmax compared to nonactivated levels and had no phosphorylation on the basis of mass spectrometry. These data suggest that MKK7 was largely responsible for JNK3 alpha 1 activation and that a single threonine phosphorylation may be all that is needed for JNK3 alpha 1 to be active. The steady-state rate constants kcat, Km(GST-ATF2++), and Km(ATP) for both monophosphorylated and bisphosphorylated JNK3 alpha 1 were within 2-fold between the two enzyme forms, suggesting the addition of tyrosine phosphorylation does not affect the binding of ATF2, ATP, or maximal turnover. Finally, the MAP kinase inhibitor, SB203580, had an IC50 value approximately 4-fold more potent on the monophosphorylated JNK3 alpha 1 compared to the bisphosphorylated JNK3 alpha 1, suggesting only a modest effect of tyrosine phosphorylation on inhibitor binding.  相似文献   

18.
Wang W  Seah SY 《Biochemistry》2005,44(27):9447-9455
HpaI, a class II pyruvate-specific aldolase involved in the catabolic pathway of hydroxyphenylacetate, is overexpressed and purified. A previous suggestion that phosphate is involved in proton transfer of pyruvate, based on the crystal structure of the homologous 2-dehydro-3-deoxygalactarate aldolase, is not substantiated from biochemical studies with HpaI. Thus, specific activities of the enzyme for the substrate 4-hydroxy-2-ketopentanoate in sodium HEPES and Tris-acetate buffers are higher than in sodium phosphate buffer. The enzyme also catalyzed the partial reaction of pyruvate proton exchange with an initial rate of 0.77 mmol min(-)(1) mg(-)(1) in phosphate-free buffer, as monitored by nuclear magnetic resonance. Steady-state kinetic analysis shows that the enzyme is also able to catalyze the aldol cleavage of 4-hydroxy-2-ketohexanoate and 3-deoxy-d-manno-oct-2-ulosonic acid (KDO). The enzyme exhibits significant oxaloacetate decarboxylase activity, with a k(cat) value 2.4-fold higher than the corresponding value for the aldol cleavage of 4-hydroxy-2-ketopentanoate. Sodium oxalate, an analogue of the enolate intermediate of the enzyme-catalyzed reaction, is a competitive inhibitor of the enzyme, with a K(i) value of 5.5 microM. Replacement of an active site arginine residue (R70) with alanine by site-specific mutagenesis resulted in an enzyme that lacks both aldolase and decarboxylase activities. The mutant enzyme is also unable to catalyze pyruvate proton exchange. The dissociation constant for pyruvate in the R70A mutant, determined by fluorescence titration, is similar to that of the wild-type enzyme, indicating that pyruvate binding is not affected by this mutation. Together, the results show that R70 influences catalysis in HpaI, particularly at the pyruvate proton exchange step.  相似文献   

19.
4-trans-(N,N-dimethylamino)cinnamaldehyde (DACA) is a chromophoric and fluorogenic substrate of aldehyde dehydrogenase. Fluorescence of DACA is enhanced by binding to aldehyde dehydrogenase in the absence of catalysis both in the presence and absence of the coenzyme analogue 5′AMP. DACA binds to aldehyde dehydrogenase with a dissociation constant of 1–3 μM and stoichiometry of 2 mol mol−1 enzyme. Incorporation of DACA during catalysis was also investigated and found to be 2 mol DACA mol−1 enzyme. Effect of pH on the stoichiometry of DACA incorporation during catalysis has shown that DACA incorporation remained constant at 2 mol DACA mol−1 enzyme, despite a 74-fold velocity enhancement between pH 5.0 and 9.0. Increase of pH increased decomposition of enzyme–acyl intermediate without affecting the rate-limiting step of the reaction. At pH 7.0 the pH stimulated velocity enhancement was 10-fold over that at pH 5.0; further velocity enhancement (11.5-fold that of pH 7.0) was achieved by 150 μM Mg2+ ions. The velocity at pH 7.0 with Mg2+ exceeded that of pH 9.0, and that at maximal pH stimulation at pH 9.5. It was observed that level of intermediate decreased to about 1 mol mol−1 enzyme, indicating that Mg2+ ions increased the rate of decomposition of the enzyme–acyl intermediate and shifted the rate-limiting step of the reaction to another step in the reaction sequence.  相似文献   

20.
Q Su  J P Klinman 《Biochemistry》1999,38(26):8572-8581
Glucose oxidase catalyzes the oxidation of glucose by molecular dioxygen, forming gluconolactone and hydrogen peroxide. A series of probes have been applied to investigate the activation of dioxygen in the oxidative half-reaction, including pH dependence, viscosity effects, 18O isotope effects, and solvent isotope effects on the kinetic parameter Vmax/Km(O2). The pH profile of Vmax/Km(O2) exhibits a pKa of 7.9 +/- 0.1, with the protonated enzyme form more reactive by 2 orders of magnitude. The effect of viscosogen on Vmax/Km(O2) reveals the surprising fact that the faster reaction at low pH (1.6 x 10(6) M-1 s-1) is actually less diffusion-controlled than the slow reaction at high pH (1.4 x 10(4) M-1 s-1); dioxygen reduction is almost fully diffusion-controlled at pH 9.8, while the extent of diffusion control decreases to 88% at pH 9.0 and 32% at pH 5.0, suggesting a transition of the first irreversible step from dioxygen binding at high pH to a later step at low pH. The puzzle is resolved by 18O isotope effects. 18(Vmax/Km) has been determined to be 1.028 +/- 0.002 at pH 5.0 and 1.027 +/- 0.001 at pH 9.0, indicating that a significant O-O bond order decrease accompanies the steps from dioxygen binding up to the first irreversible step at either pH. The results at high pH lead to an unequivocal mechanism; the rate-limiting step in Vmax/Km(O2) for the deprotonated enzyme is the first electron transfer from the reduced flavin to dioxygen, and this step accompanies binding of molecular dioxygen to the active site. In combination with the published structural data, a model is presented in which a protonated active site histidine at low pH accelerates the second-order rate constant for one electron transfer to dioxygen through electrostatic stabilization of the superoxide anion intermediate. Consistent with the proposed mechanisms for both high and low pH, solvent isotope effects indicate that proton transfer steps occur after the rate-limiting step(s). Kinetic simulations show that the model that is presented, although apparently in conflict with previous models for glucose oxidase, is in good agreement with previously published kinetic data for glucose oxidase. A role for electrostatic stabilization of the superoxide anion intermediate, as a general catalytic strategy in dioxygen-utilizing enzymes, is discussed.  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