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.
The family of FMN-dependent, alpha-hydroxy acid-oxidizing enzymes catalyzes substrate dehydrogenation by a mechanism the first step of which is abstraction of the substrate alpha-proton (so-called carbanion mechanism). For flavocytochrome b2 and lactate oxidase, it was shown that once on the enzyme this proton is lost only slowly to the solvent (Lederer F, 1984, In: Bray RC, Engel PC, Mayhew SG, eds, Flavins & flavoproteins, Berlin: Walter de Gruyter & Co., pp 513-526; Urban P, Lederer F, 1985, J Biol Chem 260:11115-11122). This suggested the occurrence of a pKa increase of the catalytic histidine upon enzyme reduction by substrate. For flavocytochrome b2, the crystal structure indicated 2 possible origins for the stabilization of the imidazolium form of His 373: either a network of hydrogen bonds involving His 373, Tyr 254, flavin N5 and O4, a heme propionate, and solvent molecules, and/or electrostatic interactions with Asp 282 and with the reduced cofactor N1 anion. In this work, we probe the effect of the hydrogen bond network at the active site by studying proton exchange with solvent for 2 mutants: Y254F and the recombinant flavodehydrogenase domain, in which this network should be disrupted. The rate of proton exchange, as determined by intermolecular hydrogen transfer experiments, appears identical in the flavodehydrogenase domain and the wild-type enzyme, whereas it is about 3-fold faster in the Y254F mutant. It thus appears that specific hydrogen bonds to the solvent do not play a major role in stabilizing the acid form of His 373 in reduced flavocytochrome b2. Removal of the Y254 phenol group induces a pKa drop of about half a pH unit for His 373 in the reduced enzyme. Even then, the rate of exchange of the imidazolium proton with solvent is still lower by several orders of magnitude than that of a normally ionizing histidine. Other factors must then also contribute to the pKa increase, such as the electrostatic interactions with D282 and the anionic reduced cofactor, as suggested by the crystal structure.  相似文献   

2.
A role for Tyr254 in L-lactate dehydrogenation catalyzed by flavocytochrome b2 has recently been proposed on the basis of the known active-site structure and of studies that had suggested a mechanism involving the initial formation of a lactate carbanion [Lederer, F., & Mathews, F.S. (1987) in Flavins and Flavoproteins, Proceedings of the Ninth International Symposium, Atlanta, GA, 1987 (Edmondson, D.E., & McCormick, D.B., Eds.) pp 133-142, Walter de Gruyter, Berlin]. This role is now examined after replacement of Tyr254 with phenylalanine. The kcat is decreased about 40-fold, Km for lactate appears unchanged, and the mainly rate-limiting step is still alpha-hydrogen abstraction, as judged from the steady-state deuterium isotope effect. Modeling studies with lactate introduced into the active site indicate two possible substrate conformations with different hydrogen-bonding partners for the substrate hydroxyl. If the hydrogen bond is formed with Tyr254, as was initially postulated, the mechanism must involve removal by His373 of the C2 hydrogen, with carbanion formation. If, in the absence of the Tyr254 phenol group, the hydrogen bond is formed with His373 N3, the substrate is positioned in such a way that the reaction must proceed by hydride transfer. Therefore the mechanism of the Y254F enzyme was investigated so as to distinguish between the two mechanistic possibilities. 2-Hydroxy-3-butynoate behaves with the mutant as a suicide reagent, as with the wild-type enzyme. Similarly, the mutant protein also catalyzes the reduction and the dehydrohalogenation of bromopyruvate under transhydrogenation conditions.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
The two preceding papers [Powers, V. M., Koo, C. W., Kenyon, G. L., Gerlt, J. A., & Kozarich, J. W. (1991) Biochemistry (first paper of three in this issue); Neidhart, D. J., Howell, P. L., Petsko, G. A., Powers, V. M., Li, R., Kenyon, G. L., & Gerlt, J. A. (1991) Biochemistry (second paper of three in this issue)] suggest that the active site of mandelate racemase (MR) contains two distinct general acid/base catalysts: Lys 166, which abstracts the alpha-proton from (S)-mandelate, and His 297, which abstracts the alpha-proton from (R)-mandelate. In this paper we report on the properties of the mutant of MR in which His 297 has been converted to asparagine by site-directed mutagenesis (H297N). The structure of H297N, solved by molecular replacement at 2.2-A resolution, reveals that no conformational alterations accompany the substitution. As expected, H297N has no detectable MR activity. However, H297N catalyzes the stereospecific elimination of bromide ion from racemic p-(bromomethyl)mandelate to give p-(methyl)-benzoylformate in 45% yield at a rate equal to that measured for wild-type enzyme; the unreacted p-(bromomethyl)mandelate is recovered as (R)-p-(hydroxymethyl)mandelate. At pD 7.5, H297N catalyzes the stereospecific exchange of the alpha-proton of (S)- but not (R)-mandelate with D2O solvent at a rate 3.3-fold less than that observed for incorporation of solvent deuterium into (S)-mandelate catalyzed by wild-type enzyme. The pD dependence of the rate of the exchange reaction catalyzed by H297N reveals a pKa of 6.4 in D2O, which is assigned to Lys 166.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

4.
The citrate transporter CitP of lactic acid bacteria catalyzes electrogenic precursor-product exchange of citrate versus L-lactate during citrate-glucose cometabolism. In the absence of sugar, L-lactate is replaced by the metabolic intermediates/end products pyruvate, α-acetolactate, and acetate. In this study, the binding and translocation properties of CitP were analyzed systematically for a wide variety of mono- and dicarboxylates of the form X-CR(2)-COO(-), where X represents OH (2-hydroxy acid), O (2-keto acid), or H (acid) and R groups differ in size, hydrophobicity, and composition. It follows that CitP is a very promiscuous carboxylate transporter. A carboxylate group is both essential and sufficient for recognition by the transporter. A C-2 atom is not essential, formate is a substrate, and C-2 may be part of a ring structure, as in benzoate. The R group may be as bulky as an indole ring structure. For all monocarboxylates of the form X-CHR-COO(-), the hydroxy (X = OH) analogs were the preferred substrates. The preference for keto (X = O) or acid (X = H) analogs was dependent on the bulkiness of the R group, such that the acid was preferred for small R groups and the 2-ketoacid was preferred for more bulky R groups. The C(4) to C(6) dicarboxylates succinate, glutarate, and adipate were also substrates of CitP. The broad substrate specificity is discussed in the context of a model of the binding site of CitP. Many of the substrates of CitP are intermediates or products of amino acid metabolism, suggesting that CitP may have a broader physiological function than its role in citrate fermentation alone.  相似文献   

5.
Bakers' yeast flavocytochrome b2 is a flavin-dependent L-2-hydroxy acid dehydrogenase which also exhibits transhydrogenase activity. When a reaction takes place between [2-3H]lactate and a halogenopyruvate, tritium is found in water and at the halogenolactate C2 position. When the halogenopyruvate undergoes halide ion elimination, tritium is also found at the C3 position of the resulting pyruvate. The amount tau of this intermolecular tritium transfer depends on the initial keto acid-acceptor concentration. At infinite acceptor concentration, extrapolation yields a maximal transfer of 97 +/- 11%. This indicates that the hydroxy acid-derived hydrogen resides transiently on enzyme monoprotic heteroatoms and that exchange with bulk solvent occurs only at the level of free reduced enzyme. Using a minimal kinetic scheme, the rate constant for hydrogen exchange between Ered and solvent is calculated to be on the order of 10(2) M-1 S-1, which leads to an estimated pK approximately equal to 15 for the ionization of the substrate-derived proton while on the enzyme. It is suggested that this hydrogen could be shared between the active site base and Flred N5 anion. It is furthermore shown that some tritium is incorporated into the products when the transhydrogenation is carried out in tritiated water. Finally, with [2-2H]lactate-reduced enzyme, a deuterium isotope effect is observed on the rate of bromopyruvate disappearance. Extrapolation to infinite bromopyruvate concentration yields DV = 4.4. An apparent inverse isotope effect is determined for bromide ion elimination. These results strengthen the idea that oxidoreduction and elimination pathways involve a common carbanionic intermediate.  相似文献   

6.
Aspartate aminotransferase (AAT) catalyzes amino group transfer from glutamate (Glu) or aspartate (Asp) to a keto acid acceptor-oxaloacetate (OA) or alpha-ketoglutarate (KG), respectively. Data presented here show that AAT catalyzes two partial reactions resulting in isotope exchange between 3H-labeled Glu or 3H-labeled Asp and the cognate keto acid in the absence of the keto acid acceptor required for the net reaction. Tritiated keto acid product was detected by release of 3H2O from C-3 during base-induced enolization. Tritium released directly from C-2 (or C-3) by the enzyme was also evaluated and is a small fraction of that released because of exchange to the keto acid pool. Exchange is dependent on AAT concentration, time-dependent, proportional to the amino-to-keto acid ratio, and blocked by aminooxyacetate (AOA), an AAT inhibitor. Enzymatic conversion of [3H]KG to Glu by glutamic dehydrogenase (GDH) or of [3H]OA to malate by malic dehydrogenase (MDH) "protects" the label from release by base, showing that base-induced isotope release is from keto acid rather than a result of release during the exchange process. AAT isotope exchange is discussed in the context of the glutamate/glutamine shuttle hypothesis for astrocyte/neuron carbon cycling.  相似文献   

7.
Flavocytochrome b2 (or L-lactate dehydrogenase) from baker's yeast is thought to operate by the initial formation of a carbanion, as do the evolutionarily related alpha-hydroxy acid-oxidizing FMN-dependent oxidases. Previous work has shown that, in the active site of the unligated reduced flavocytochrome b2, the group that has captured the substrate alpha-proton has a high pKapp, calculated to lie around 15 through the use of Eigen's equation. A detailed inspection of the now known three-dimensional structure of the enzyme leads to the conclusion that the high pKa belongs to His 373, an active site group that plays the role of general base in the forward reaction and of general acid in the reverse direction. Moreover, consideration of the kinetics of proton transfer during the catalytic cycle suggests that the pKa of the reduced FMN N5 position should be lowered by several pH units compared to its pKa of 20 or more when free. The features of the three-dimensional structure possibly responsible for these pK shifts are analyzed; they are proposed to consist of a network of hydrogen bonds with the solvent and of a mutual electrostatic stabilization of anionic reduced flavin and the imidazolium ion. Finally, it is suggested that similar pK shifts affect the active sites of the alpha-hydroxy acid-oxidizing flavooxidases, which are homologous to flavocytochrome b2. The functional significance of these pK shifts in terms of catalysis and semiquinone stabilization is discussed.  相似文献   

8.
The flavoprotein oxidase Fms1 from Saccharomyces cerevisiae catalyzes the oxidation of spermine and N(1)-acetylspermine to spermidine and 3-aminopropanal or N-acetyl-3-aminopropanal. Within the active site of Fms1, His67 is positioned to form hydrogen bonds with the polyamine substrate. This residue is also conserved in other polyamine oxidases. The catalytic properties of H67Q, H67N, and H67A Fms1 have been characterized to evaluate the role of this residue in catalysis. With both spermine and N(1)-acetylspermine as the amine substrate, the value of the first-order rate constant for flavin reduction decreases 2-3 orders of magnitude, with the H67Q mutation having the smallest effect and H67N the largest. The k(cat)/K(O2) value changes very little upon mutation with N(1)-acetylspermine as the amine substrate and decreases only an order of magnitude with spermine. The k(cat)/K(M)-pH profiles with N(1)-acetylspermine are bell-shaped for all the mutants; the similarity to the profile of the wild-type enzyme rules out His67 as being responsible for either of the pK(a) values. The pH profiles for the rate constant for flavin reduction for all the mutant enzymes similarly show the same pK(a) as wild-type Fms1, about ~7.4; this pK(a) is assigned to the substrate N4. The k(cat)/K(O2)-pH profiles for wild-type Fms1 and the H67A enzyme both show a pK(a) of about ~6.9; this suggests His67 is not responsible for this pH behavior. With the H67Q, H67N, and H67A enzymes the k(cat) value decreases when a single residue is protonated, as is the case with the wild-type enzyme. The structure of H67Q Fms1 has been determined at a resolution of 2.4 ?. The structure shows that the mutation disrupts a hydrogen bond network in the active site, suggesting that His67 is important both for direct interactions with the substrate and to maintain the overall active site structure.  相似文献   

9.
S Ghisla  C Thorpe  V Massey 《Biochemistry》1984,23(14):3154-3161
Butyryl-CoA dehydrogenase from Megasphera elsdenii catalyzes the exchange of the alpha- and beta-hydrogens of substrate with solvent [Gomes, B., Fendrich, G., & Abeles, R. H. (1981) Biochemistry 20, 1481-1490]. The stoichiometry of this exchange was determined by using 3H2O label as 1.94 +/- 0.1 per substrate molecule. The rate of 3H label incorporation into substrate under anaerobic conditions is monophasic, indicating that both the alpha- and beta-hydrogens exchange at the same rate. The exchange in 2H2O leads to incorporation of one 2H each into the alpha- and the beta-positions of butyryl-CoA, as determined by companion 1H NMR experiments and confirmed by mass spectroscopic analysis. In contrast, with general acyl-CoA dehydrogenase from pig kidney, only exchange of the alpha-hydrogen was found. The beta-hydrogen is the one that is transferred (reversibly) to the flavin 5-position during substrate dehydrogenation. This was demonstrated by reacting 5-3H- and 5-2H-reduced 5-deaza-FAD-general acyl-CoA dehydrogenase with crotonyl-CoA. Only one face of the reduced flavin analogue is capable of transferring hydrogen to substrate. The rate of this reaction is 11.1 s-1 for 5-deaza-FAD-enzyme and 2.2 s-1 for [5-2H]deaza-FAD-enzyme, yielding an isotope effect of 5. These values compare with a rate of 2.6 s-1 for the reaction of native reduced enzyme with crotonyl-CoA. The two reduced enzymes (normal vs. 5-deaza-FAD-enzyme) thus react at similar rates, indicating a similar mechanism.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
Williams L  Nguyen T  Li Y  Porter TN  Raushel FM 《Biochemistry》2006,45(24):7453-7462
Uronate isomerase, a member of the amidohydrolase superfamily, catalyzes the isomerization of D-glucuronate and D-fructuronate. During the interconversion of substrate and product the hydrogen at C2 of D-glucuronate is transferred to the pro-R position at C1 of the product, D-fructuronate. The exchange of the transferred hydrogen with solvent deuterium occurs at a rate that is 4 orders of magnitude slower than the interconversion of substrate and product. The enzyme catalyzes the elimination of fluoride from 3-deoxy-3-fluoro-D-glucuronate. These results have been interpreted to suggest a chemical reaction mechanism in which an active site base abstracts the proton from C2 of D-glucuronate to form a cis-enediol intermediate. The conjugate acid then transfers this proton to C1 of the cis-enediol intermediate to form D-fructuronate. The loss of fluoride from 3-deoxy-3-fluoro-D-glucuronate is consistent with a stabilized carbanion at C2 of the substrate during substrate turnover. The slow exchange of the transferred hydrogen with solvent water is consistent with a shielded conjugate acid after abstraction of the proton from either D-glucuronate or D-fructuronate during the isomerization reaction. This conclusion is supported by the competitive inhibition of the enzymatic reaction by D-arabinaric acid and the monohydroxamate derivative with Ki values of 13 and 670 nM, respectively. There is no evidence to support a hydride transfer mechanism for uronate isomerase. The wild type enzyme was found to contain 1 equiv of zinc per subunit. The divalent cation could be removed by dialysis against the metal chelator, dipicolinate. However, the apoenzyme has the same catalytic activity as the Zn-substituted enzyme and thus the divalent metal ion is not required for enzymatic activity. This is the only documented example of a member in the amidohydrolase superfamily that does not require one or two divalent cations for enzymatic activity.  相似文献   

11.
Flavocytochrome b2 catalyses the oxidation of L-lactate to pyruvate in yeast mitochondrial intermembrane space. Its flavoprotein domain is a member of a family of FMN-dependent 2-hydroxy-acid-oxidizing enzymes. Numerous solution studies suggest that the first step of the reaction consists of proton abstraction from lactate C2, leading to a carbanion that subsequently yields electrons to FMN. The crystal structure suggests that the enzyme base is His373, and that Tyr254 may be hydrogen bonded to the substrate hydroxyl. Studies carried out with the Y254F mutant [Dubois, J., Chapman, S.K., Mathews, F.S., Reid, G.A. & Lederer, F. (1990) Biochemistry 29, 6393-6400] showed that Tyr254 does not act as a base but stabilizes the transition state. As the mutation did not induce any change in substrate affinity, the question of the existence of the hydrogen bond in the Michaelis complex remained open. Similar results with glycolate oxidase, mutated at the same position, led to the suggestion that these enzymes actually operate via a hydride transfer mechanism [Macheroux, P., Kieweg, V., Massey, V., Soderlind, E., Stenberg, K. & Lindqvist, Y. (1993) Eur. J. Biochem. 213, 1047-1054]. In the present work, we have re-investigated the matter by analysing the properties of a Y254L mutant flavocytochrome b2, as well as the behaviour of the Y254F enzyme with two substrates other than lactate, and a series of inhibitors. The Y254L protein is less efficient with L-lactate than the wild-type enzyme by a factor of 500, but the substrate affinity is unchanged. In contrast, L-phenyllactate and mandelate, poor substrates (the latter acting more as an inhibitor), exhibit an increased affinity. In addition, the Y254L mutant enzyme is more efficient with phenyllactate than lactate as a substrate. In order to rationalize these observations, we have modelled phenyllactate and mandelate in the active site, using previously described modelling experiments with lactate as a starting point. The results indicate that mandelate cannot bind in an orientation allowing proton abstraction by His373, due to steric interference by the side chains of Ala198 and Leu230. It might possibly adopt a binding mode as proposed previously for lactate, which leads to a hydride transfer and with which the 198 and 230 side chains do not interfere. However, other researchers [Sinclair, R., Reid, G.A. & Chapman, S.K. (1998) Biochem. J. 333, 117-120] showed that A198G, L230A and A198G/L230A mutant enzymes exhibit a strongly improved mandelate dehydrogenase activity. These results indicate that relief of the steric crowding facilitates catalysis by enabling a better mandelate orientation at the active site, suggesting that its productive binding mode is similar to that proposed for lactate in the carbanion mechanism. The modelling studies therefore support the hypothesis of a carbanion mechanism for all substrates. In addition, we present the effect of the two mutations at position 254 on the binding of a number of competitive inhibitors (such as sulfite, D-lactate, propionate) and of inhibitors that are known to bind at the active site both when the flavin is oxidized and when it is in the semiquinone state (propionate, oxalate and L-lactate at high concentrations). Unexpectedly, the results indicate that the integrity of Tyr254 is necessary for the binding of these inhibitors at the semiquinone stage.  相似文献   

12.
Aminodeoxychorismate lyase is a pyridoxal 5'-phosphate-dependent enzyme that converts 4-aminodeoxychorismate to pyruvate and p-aminobenzoate, a precursor of folic acid in bacteria. The enzyme exhibits significant sequence similarity to two aminotransferases, D-amino acid aminotransferase and branched-chain L-amino acid aminotransferase. In the present study, we have found that aminodeoxychorismate lyase catalyzes the transamination between D-alanine and pyridoxal phosphate to produce pyruvate and pyridoxamine phosphate. L-Alanine and other D- and L-amino acids tested were inert as substrates of transamination. The pro-R hydrogen of C4' of pyridoxamine phosphate was stereospecifically abstracted during the reverse half transamination from pyridoxamine phosphate to pyruvate. Aminodeoxychorismate lyase is identical to D-amino acid aminotransferase and branched-chain L-amino acid aminotransferase in the stereospecificity of the hydrogen abstraction, and differs from all other pyridoxal enzymes that catalyze pro-S hydrogen transfer. Aminodeoxychorismate lyase is the first example of a lyase that catalyzes pro-R-specific hydrogen abstraction. The result is consistent with recent X-ray crystallographic findings showing that the topological relationships between the cofactor and the catalytic residue for hydrogen abstraction are conserved among aminodeoxychorismate lyase, D-amino acid aminotransferase and branched-chain L-amino acid aminotransferase [Nakai, T., Mizutani, H., Miyahara, I., Hirotsu, K., Takeda, S., Jhee, K.-H., Yoshimura, T., and Esaki, N. (2000) J. Biochem. 128, 29-38].  相似文献   

13.
BphD of Burkholderia xenovorans LB400 catalyzes an unusual C-C bond hydrolysis of 2-hydroxy-6-oxo-6-phenylhexa-2,4-dienoic acid (HOPDA) to afford benzoic acid and 2-hydroxy-2,4-pentadienoic acid (HPD). An enol-keto tautomerization has been proposed to precede hydrolysis via a gem-diol intermediate. The role of the canonical catalytic triad (Ser-112, His-265, Asp-237) in mediating these two half-reactions remains unclear. We previously reported that the BphD-catalyzed hydrolysis of HOPDA (lambda(max) is 434 nm for the free enolate) proceeds via an unidentified intermediate with a red-shifted absorption spectrum (lambda(max) is 492 nm) (Horsman, G. P., Ke, J., Dai, S., Seah, S. Y. K., Bolin, J. T., and Eltis, L. D. (2006) Biochemistry 45, 11071-11086). Here we demonstrate that the S112A variant generates and traps a similar intermediate (lambda(max) is 506 nm) with a similar rate, 1/tau approximately 500 s(-1). The crystal structure of the S112A:HOPDA complex at 1.8-A resolution identified this intermediate as the keto tautomer, (E)-2,6-dioxo-6-phenyl-hex-3-enoate. This keto tautomer did not accumulate in either the H265A or the S112A/H265A double variants, indicating that His-265 catalyzes tautomerization. Consistent with this role, the wild type and S112A enzymes catalyzed tautomerization of the product HPD, whereas H265A variants did not. This study thus identifies a keto intermediate, and demonstrates that the catalytic triad histidine catalyzes the tautomerization half-reaction, expanding the role of this residue from its purely hydrolytic function in other serine hydrolases. Finally, the S112A:HOPDA crystal structure is more consistent with hydrolysis occurring via an acyl-enzyme intermediate than a gem-diol intermediate as solvent molecules have poor access to C6, and the closest ordered water is 7 A away.  相似文献   

14.
Biosynthetic origin of hydrogen atoms in the lipase inhibitor lipstatin   总被引:1,自引:0,他引:1  
The lipase inhibitor lipstatin is biosynthesized in Streptomyces toxytricini via condensation of a C(14) precursor and a C(8) precursor, which are both obtained from fatty acid catabolism. To study the mechanism of this reaction in more detail, S. toxytricini was grown in medium containing a mixture of U-(13)C,U-(2)H-lipids and unlabeled sunflower oil or in a medium containing 70% D(2)O. Lipstatin was isolated and analyzed by (1)H,(2)H, and (13)C NMR spectroscopy. Hydrogen atoms at C-2, C-3, and C-4 of lipstatin were found to be derived from solvent protons. The formation of the lipstatin precursor 3-hydroxy-Delta(5,8)-tetradecadienoyl-CoA by beta oxidation of linoleic acid explains the incorporation of solvent hydrogen into the 4 position of lipstatin. The hydrogen in position 3 of lipstatin is most probably introduced from solvent by proton/deuterium exchange of a redox cofactor involved in the reduction of the keto group in the branched chain beta keto acid arising by a decarboxylative condensation. The incorporation of solvent hydrogen at position 2 can be explained by epimerization of a chiral intermediate at C-2 and C-3. Epimerization may involve a dehydration-rehydration mechanism.  相似文献   

15.
In the course of a half-reaction of enzymic transamination, the aldimine adduct formed between the coenzyme pyridoxal 5'-phosphate and the amino acid substrate tautomerizes to the ketimine intermediate which is then hydrolyzed to the oxo acid product and the pyridoxamine 5'-phosphate form of the enzyme. In the reverse half-reaction the tautomerization is initiated by the removal of a proton from the pro-S position at C-4' of the PMP moiety of the ketimine intermediate. The present study investigates the question whether the pro-S hydrogen at C-4' of PMP is labilized by its active site environment independently of the formation of the ketimine intermediate, i.e. in the absence of substrate. Reconstitution of apoaspartate aminotransferase (mitochondrial isoenzyme from chicken) with [4'-3H] PMP results indeed in a stereospecific exchange of pro-S 3H with solvent water. The exchange follows first order kinetics (t 1/2 = 23 min at pH 7.5 and 25 degrees C). Unbound PMP showed no measurable exchange. Rigorous control experiments excluded the possibility that the observed exchange was due to a transamination reaction of the enzyme with contaminating oxo acid substrates. The newly observed stereospecific exchange reaction allows to investigate the acid/base properties of C-4' and the modulating effects of its active site environment independently of the preceding and following steps of enzymic transamination.  相似文献   

16.
Yue QK  Kass IJ  Sampson NS  Vrielink A 《Biochemistry》1999,38(14):4277-4286
Cholesterol oxidase is a monomeric flavoenzyme which catalyzes the oxidation and isomerization of cholesterol to cholest-4-en-3-one. The enzyme interacts with lipid bilayers in order to bind its steroid substrate. The X-ray structure of the enzyme from Brevibacterium sterolicum revealed two loops, comprising residues 78-87 and residues 433-436, which act as a lid over the active site and facilitate binding of the substrate [Vrielink et al. (1991) J. Mol. Biol. 219, 533-554; Li et al. (1993) Biochemistry 32, 11507-11515]. It was postulated that these loops must open, forming a hydrophobic channel between the membrane and the active site of the protein and thus sequestering the cholesterol substrate from the aqueous environment. Here we describe the three-dimensional structure of the homologous enzyme from Streptomyces refined to 1.5 A resolution. Structural comparisons to the enzyme from B. sterolicum reveal significant conformational differences in these loop regions; in particular, a region of the loop comprising residues 78-87 adopts a small amphipathic helical turn with hydrophobic residues directed toward the active site cavity and hydrophilic residues directed toward the external surface of the molecule. It seems reasonable that this increased rigidity reduces the entropy loss that occurs upon binding substrate. Consequently, the Streptomyces enzyme is a more efficient catalyst. In addition, we have determined the structures of three active site mutants which have significantly reduced activity for either the oxidation (His447Asn and His447Gln) or the isomerization (Glu361Gln). Our structural and kinetic data indicate that His447 and Glu361 act as general base catalysts in association with conserved water H2O541 and Asn485. The His447, Glu361, H2O541, and Asn485 hydrogen bond network is conserved among other oxidoreductases. This catalytic tetrad appears to be a structural motif that occurs in flavoenzymes that catalyze the oxidation of unactivated alcohols.  相似文献   

17.
Chorismate synthase catalyzes the last step in the common shikimate pathway leading to aromatic compounds such as the aromatic amino acids. The reaction consists of the 1,4-anti-elimination of the 3-phosphate group and the C-(6proR) hydrogen from 5-enolpyruvylshikimate 3-phosphate to yield chorismate. Although this reaction does not involve a net redox change, the enzyme has an absolute requirement for reduced flavin mononucleotide, which is not consumed during the reaction. Two invariant histidine residues are found in the active site of the enzyme: His(17) and His(106). Using site-directed mutagenesis, both histidines were replaced by alanine, reducing the activity 10- and 20-fold in the H106A and H17A mutant protein, respectively. Based on the characterization of the two single mutant proteins, it is proposed that His(106) serves to protonate the monoanionic reduced FMN, whereas His(17) protonates the leaving phosphate group of the substrate. An enzymatic reaction mechanism in keeping with the experimental results is presented.  相似文献   

18.
Initial velocity, product inhibition, and substrate inhibition studies suggest that the endogenous lactate dehydrogenase activity of duck epsilon-crystallin follows an order Bi-Bi sequential mechanism. In the forward reaction (pyruvate reduction), substrate inhibition by pyruvate was uncompetitive with inhibition constant of 6.7 +/- 1.7 mM. In the reverse reaction (lactate oxidation), substrate inhibition by L-lactate was uncompetitive with inhibition constant of 158 +/- 25 mM. The cause of these inhibitions may be due to epsilon-crystallin-NAD(+)-pyruvate and epsilon-crystallin-NADH-L-lactate abortive ternary complex formation as suggested by the multiple inhibition studies. Pyruvate binds to free enzyme very poorly, with a very large dissociation constant. Bromopyruvate, fluoropyruvate, pyruvate methyl ester, and pyruvate ethyl ester are alternative substrates for pyruvate. 3-Acetylpyridine adenine dinucleotide, nicotinamide 1,N6-ethenoadenine dinucleotide, and nicotinamide hypoxanthine dinucleotide serve as alternative coenzymes for epsilon-crystallin. All the above alternative substrates or coenzymes showed an intersecting initial-velocity pattern conforming to the order Bi--Bi kinetic mechanism. Nicotinic acid adenine dinucleotide, thionicotinamide adenine dinucleotide, and 3-aminopyridine adenine dinucleotide acted as inhibitors for this enzymatic crystallin. The inhibitors were competitive versus NAD+ and noncompetitive versus L-lactate. alpha-NAD+ was a noncompetitive inhibitor with respect to the usual beta-NAD+. D-Lactate, tartronate, and oxamate were strong dead-end inhibitors for the lactate dehydrogenase activity of epsilon-crystallin. Both D-lactate and tartronate were competitive inhibitors versus L-lactate while oxamate was a competitive inhibitor versus pyruvate. We conclude that the structural requirements for the substrate and coenzyme of epsilon-crystallin are similar to those of other dehydrogenases and that the carboxamide carbonyl group of the nicotinamide moiety is important for the coenzyme activity.  相似文献   

19.
Aromatase (CYP19) catalyzes three consecutive hydroxylation reactions converting C19 androgens to aromatic C18 estrogenic steroids. In this study, five human aromatase mutants (E302D, S478A, S478T, H480K, and H480Q) were prepared using a mammalian cell expression system. These mutants were evaluated by enzyme kinetic analysis, inhibitory profile studies, and reaction intermediate measurements. Three steroidal inhibitors [4-hydroxyandrostenedione (4-OHA), 7alpha-(4'-amino)phenylthio-1,4-androstandiene-3,17-dione (7alpha-APTADD), and bridge (2,19-methyleneoxy) androstene-3,17-dione (MDL 101003)], and four nonsteroidal inhibitors [aminoglutethimide (AG), CGS 20267, ICI D1033, and vorozole (R83842)] were used in the inhibitory profile studies. Our computer model of aromatase suggests that Glu302 is situated in the conserved I-helix region and located near the C-19 position of the steroid substrate. The model was supported by significant changes in kinetic parameters and a sevenfold increase in the Ki value of MDL 101,003 for the mutant E302D. As S478A was found to have kinetic properties similar to the wild-type enzyme and a much higher activity than S478T, Ser478 is thought to be situated in a rather restricted environment. There was a 10-fold increase in the Ki value of 7alpha-APTADD for S478T over that for the wild-type enzyme, suggesting that Ser478 might be near the C-7 position of the substrate. The reaction intermediate analysis revealed that significantly more 19-ol intermediate was generated by both S478A and S478T than the wild-type enzyme. These results would support a hypothesis that Ser478 plays a role in the first and second hydroxylation reactions. A positive charged amino acid is preferred at position 480 as shown by the fact that H480K has a significantly higher activity than H480Q. The Ki value of 4-OHA for H480Q was found to be three times that of the wild-type enzyme. In addition, significantly more 19-ol and 19-al intermediates were detected for both mutants H480K and H480Q than for the wild-type enzyme. Evaluation of the two mutations at His480 allows us to propose that this residue may participate in the aromatization reaction (the third step) by acting as a hydrogen bond donor for the C-3 keto group of the substrate. Furthermore, new products were generated when the enzyme was mutated at Ser478 and His480. Thus, these two residues must play an important role in the catalysis and are likely closer to the substrate binding site than previously predicted.  相似文献   

20.
The crotonase homolog, 6-oxo camphor hydrolase (OCH), catalyzes the desymmetrization of bicyclic beta-diketones to optically active keto acids via an enzymatic retro-Claisen reaction, resulting in the cleavage of a carbon-carbon bond. We have previously reported the structure of OCH (Whittingham, J. L., Turkenburg, J. P., Verma, C. S., Walsh, M. A., and Grogan, G. (2003) J. Biol. Chem. 278, 1744-1750), which suggested the involvement of five residues, His-45, His-122, His-145, Asp-154, and Glu-244, in catalysis. Here we report mutation studies on OCH that reveal that H145A and D154N mutants of OCH have greatly reduced values of k(cat)/K(m) derived from a very large increase in K(m) for the native substrate, 6-oxo camphor. In addition, H122A has a greatly reduced value of k(cat), and its K(m) is five times that of the wild-type. The location of the active site is confirmed by the 1.9-A structure of the H122A mutant of OCH complexed with the minor diastereoisomer of (2S,4S)-alpha-campholinic acid, the natural product of the enzyme. This shows the pendant acetate of the product hydrogen bonded to a His-145/Asp-154 dyad and the endocyclic carbonyl of the cyclopentane ring hydrogen bonded to Trp-40. The results are suggestive of a base-catalyzed mechanism of C-C bond cleavage and provide clues to the origin of prochiral selectivity by the enzyme and to the recruitment of the crotonase fold for alternate modes of transition state stabilization to those described for other crotonase superfamily members.  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