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 E. coli isozyme of gamma-aminobutyrate aminotransferase (GABA-AT) is a tetrameric pyridoxal phosphate-dependent enzyme that catalyzes transamination between primary amines and alpha-keto acids. The roles of the active site residues V241, E211, and I50 in the GABA-AT mechanism have been probed by site-directed mutagenesis. The beta-branched side chain of V241 facilitates formation of external aldimine intermediates with primary amine substrates, while E211 provides charge compensation of R398 selectively in the primary amine half-reaction and I50 forms a hydrophobic lid at the top of the substrate binding site. The structures of the I50Q, V241A, and E211S mutants were solved by X-ray crystallography to resolutions of 2.1, 2.5, and 2.52 A, respectively. The structure of GABA-AT is similar in overall fold and active site structure to that of dialkylglycine decarboxylase, which catalyzes both transamination and decarboxylation half-reactions in its normal catalytic cycle. Therefore, an attempt was made to convert GABA-AT into a decarboxylation-dependent aminotransferase similar to dialkylglycine decarboxylase by systematic mutation of E. coli GABA-AT active site residues. Two of the twelve mutants presented, E211S/I50G/C77K and E211S/I50H/V80D, have approximately 10-fold higher decarboxylation activities than the wild-type enzyme, and the E211S/I50H/V80D has formally changed the reaction specificity to that of a decarboxylase.  相似文献   

2.
A homology model for the pig isozyme of the pyridoxal phosphate-dependent enzyme gamma-aminobutyrate (GABA) aminotransferase has been built based mainly on the structure of dialkylglycine decarboxylase and on a multiple sequence alignment of 28 evolutionarily related enzymes. The proposed active site structure is presented and analyzed. Hypothetical structures for external aldimine intermediates explain several characteristics of the enzyme. In the GABA external aldimine model, the pro-S proton at C4 of GABA, which abstracted in the 1,3-azaallylic rearrangement interconverting the aldimine and ketimine intermediates, is oriented perpendicular to the plane of the pyridoxal phosphate ring. Lys 329 is in close proximity and is probably the general base catalyst for the proton transfer reaction. The carboxylate group of GABA interacts with Arg 192 and Lys 203, which determine the specificity of the enzyme for monocarboxylic omega-amino acids such as GABA. In the proposed structure for the L-glutamate external aldimine, the alpha-carboxylate interacts with Arg 445. Glu 265 is proposed to interact with this same arginine in the GABA external aldimine, enabling the enzyme to act on omega-amino acids in one half-reaction and on alpha-amino acids in the other. The reactivities of inhibitors are well explained by the proposed active site structure. The R and S isomers of beta-substituted phenyl and p-chlorophenyl GABA would bind in very different modes due to differential steric interactions, with the reactive S isomer leaving the orientation of the GABA moiety relatively unperturbed compared to that of the natural substrate. In our model, only the reactive S isomer of the mechanism-based inhibitor vinyl-GABA, an effective anti-epileptic drug known clinically as Vigabatrin, would orient the scissile C4-H bond perpendicular to the coenzyme ring plane and present the proton to Lys 329, the proposed general base catalyst of the reaction. The R isomer would direct the vinyl group toward Lys 329 and the C4-H bond toward Arg 445. The active site model presented provides a basis for site-directed mutagenesis and drug design experiments.  相似文献   

3.
The crystal structures of four inhibitor complexes of dialkylglycine decarboxylase are reported. The enzyme does not undergo a domain closure, as does aspartate aminotransferase, upon inhibitor binding. Two active-site conformations have been observed in previous structures that differ in alkali metal ion content, and two active-site conformations have been shown to coexist in solution when a single type of metal ion is present. There is no indication of coexisting conformers in the structures reported here or in the previously reported structures, and the observed conformation is that expected based on the presence of potassium in the enzyme. Thus, although two active-site conformations coexist in solution, a single conformation, corresponding to the more active enzyme, predominates in the crystal. The structure of 1-aminocyclopropane-1-carboxylate bound in the active site shows the aldimine double bond to the pyridoxal phosphate cofactor to be fully out of the plane of the coenzyme ring, whereas the Calpha-CO2(-) bond lies close to it. This provides an explanation for the observed lack of decarboxylation reactivity with this amino acid. The carboxylate groups of both 1-aminocyclopropane-1-carboxylate and 5'-phosphopyridoxyl-2-methylalanine interact with Ser215 and Arg406 as previously proposed. This demonstrates structurally that alternative binding modes, which constitute substrate inhibition, occur in the decarboxylation half-reaction. The structures of d and l-cycloserine bound to the active-site show that the l-isomer is deprotonated at C(alpha), presumably by Lys272, while the d-isomer is not. This difference explains the approximately 3000-fold greater potency of the l versus the d-isomer as a competitive inhibitor of dialkylglycine decarboxylase.  相似文献   

4.
The first functional catalytic mimic of the enzyme dialkylglycine decarboxylase is described. This system utilizes a hydrophobically modified polyethylenimine polymer, a pyridoxamine cofactor, and a 2-aryl-2-alkylglycine sacrificial amine source to convert alpha-keto acids to alpha-amino acids at biologically relevant temperatures with multiple turnovers of the pyridoxamine catalyst. The effects of hydrophobic and electronic factors in the 2,2-disubstituted sacrificial amine source and the pyridoxamine catalyst on turnover frequency and turnover number are explored.  相似文献   

5.
The three-dimensional structures of pyridoxal 5'-phosphate-type aspartate aminotransferase (AspAT) from Thermus thermophilus HB8 and pyridoxamine 5'-phosphate type one in complex with maleate have been determined by X-ray crystallography at 1.8 and 2.6 A resolution, respectively. The enzyme is a homodimer, and the polypeptide chain of the subunit is folded into one arm, one small domain, and one large domain. AspATs from many species were classified into aminotransferase subgroups Ia and Ib. The enzyme belongs to subgroup Ib, its sequence being less than 16% identical to the primary sequences of Escherichia coli, pig cytosolic, and chicken mitochondrial AspATs, which belong to subgroup Ia whose sequences are more than 40% identical and whose three-dimensional structures are quite similar with the active site residues almost completely conserved. The first X-ray analysis of AspAT subgroup Ib indicated that the overall and the active site structures are essentially conserved between the AspATs of subgroup Ia and the enzyme of subgroup Ib, but there are two distinct differences between them. (1) In AspAT subgroup Ia, substrate (or inhibitor) binding induces a large movement of the small domain as a whole to close the active site. However, in the enzyme of subgroup Ib, only the N-terminal region (Lys13-Val30) of the small domain approaches the active site to interact with the maleate. (2) In AspAT subgroup Ia, Arg292 recognizes the side chain carboxylate of the substrate; however, residue 292 of the enzyme in subgroup Ib is not Arg, and in place of Arg292, Lys109 forms a salt bridge with the side chain carboxylate. The thermostability of the enzyme is attained at least in part by the high content of Pro residues in the beta-turns and the marked increase in the number of salt bridges on the molecular surface compared with the mesophilic AspAT.  相似文献   

6.
The vitamin B(6)-dependent enzyme 7,8-diaminopelargonic acid (DAPA) synthase catalyzes the antepenultimate step in the synthesis of biotin, the transfer of the alpha-amino group of S-adenosyl-l-methionine (SAM) to 7-keto-8-aminopelargonic acid (KAPA) to form DAPA. The Y17F, Y144F, and D147N mutations in the active site were constructed independently. The k(max)/K(m)(app) values for the half-reaction with DAPA of the Y17F and Y144F mutants are reduced by 1300- and 2900-fold, respectively, compared to the WT enzyme. Crystallographic analyses of these mutants do not show significant changes in the structure of the active site. The kinetic deficiencies, together with a structural model of the enzyme-PLP/DAPA Michaelis complex, point to a role of these two residues in recognition of the DAPA/KAPA substrates and in catalysis. The k(max)/K(m)(app) values for the half-reaction with SAM are similar to that of the WT enzyme, showing that the two tyrosine residues are not involved in this half-reaction. Mutations of the conserved Arg253 uniquely affect the SAM kinetics, thus establishing this position as part of the SAM binding site. The D147N mutant is catalytically inactive in both half-reactions. The structure of this mutant exhibits significant changes in the active site, indicating that this residue plays an important structural role. Of the four residues examined, only Tyr144 and Arg253 are strictly conserved in the available amino acid sequences of DAPA synthases. This enzyme thus provides an illustrative example that active site residues essential for catalysis are not necessarily conserved, i.e., that during evolution alternative solutions for efficient catalysis by the same enzyme arose. Decarboxylated SAM [S-adenosyl-(5')-3-methylthiopropylamine] reacts nearly as well as SAM and cannot be eliminated as a putative in vivo amino donor.  相似文献   

7.
Lys-145 of the thermostable D-amino acid aminotransferase, which binds pyridoxal phosphate, was replaced by Ala or Arg by site-directed mutagenesis. Both mutant enzymes were purified to homogeneity; their absorption spectra indicated that both mutant enzymes contained pyridoxal phosphate bound non-covalently. Even though the standard assay method did not indicate any activity with either mutant, addition of an amino donor, D-alanine, to the Arg-145 mutant enzyme led to a slow decrease in absorption at 392 nm with a concomitant increase in absorption at 333 nm. This result suggests that the enzyme was converted into the pyridoxamine phosphate form. The amount of pyruvate formed was almost equivalent to that of the reactive pyridoxal phosphate in the mutant enzyme. Thus, the Arg-145 mutant enzyme is able to catalyze slowly the half-reaction of transamination. Exogenous amines, such as methylamine, had no effect on the half-reaction with the Arg-145 mutant enzyme. In contrast, the Ala-145 mutant enzyme neither underwent the spectral change by addition of D-alanine nor catalyzed pyruvate formation, in the absence of added amine. However, the Ala-145 mutant enzyme catalyzed the half-reaction significantly in the presence of added amine. These findings suggest that a basic amino acid residue, such as lysine or arginine, is required at position 145 for catalysis of the half-reaction. The role of the exogenous amines differs with various active-site mutant enzymes.  相似文献   

8.
The complete amino acid sequence of bacterial omega-amino acid:pyruvate aminotransferase (omega-APT) was determined from its primary structure. The enzyme protein was fragmented by CNBr cleavage, trypsin, and Staphylococcus aureus V8 digestions. The peptides were purified and sequenced by Edman degradation. omega-ATP is composed of four identical subunits of 449 amino acids each. The calculated molecular weight of the enzyme subunit is 48,738 and that of the enzyme tetramer is 194,952. No disulfide bonds or bound sugar molecules were found in the enzyme structure, although 6 cysteine residues were determined per enzyme subunit. Sequence homologies were found between an omega-aminotransferase, i.e. mammalian and yeast ornithine delta-aminotransferases, fungal gamma-aminobutyrate aminotransferase and 7,8-diaminoperalgonate aminotransferase, and 2,2-dialkylglycine decarboxylase. The enzyme structure is not homologous to those of aspartate aminotransferases (AspATs) including the enzymes of Escherichia coli and Sufolobus salfactaricus, though significant homology in the three-dimensional structures around the cofactor binding site has been found between omega-APT and AspATs (Watanabe, N., Sakabe, K., Sakabe, N., Higashi, T., Sasaki, K., Aibara, S., Morita, Y., Yonaha, K., Toyama, S., and Fukutani, H. (1989) J. Biochem. 105, 1-3).  相似文献   

9.
The flavoprotein nitroalkane oxidase (NAO) catalyzes the oxidation of primary and secondary nitroalkanes to the corresponding aldehydes and ketones. The enzyme is a homologue of acyl-CoA dehydrogenase. Asp402 in NAO has been proposed to be the active site base responsible for removing the substrate proton in the first catalytic step; structurally it corresponds to the glutamate which acts as the base in medium chain acyl-CoA dehydrogenase. In the active site of NAO, the carboxylate of Asp402 forms an ionic interaction with the side chain of Arg409. The R409K enzyme has now been characterized kinetically and structurally. The mutation results in a decrease in the rate constant for proton abstraction of 100-fold. Analysis of the three-dimensional structure of the R409K enzyme, determined by X-ray crystallography to a resolution of 2.65 A, shows that the critical structural change is an increase in the distance between the carboxylate of Asp402 and the positively charged nitrogen in the side chain of the residue at position 409. The D402E mutation results in a smaller decrease in the rate constant for proton abstraction of 18-fold. The structure of the D402E enzyme, determined at 2.4 A resolution, shows that there is a smaller increase in the distance between Arg409 and the carboxylate at position 402, and the interaction of this residue with Ser276 is perturbed. These results establish the critical importance of the interaction between Asp402 and Arg409 for proton abstraction by nitroalkane oxidase.  相似文献   

10.
Aromatic amino acid aminotransferase is active toward both aromatic and dicarboxylic amino acids, and the mechanism for this dual substrate recognition has been an issue in the enzymology of this enzyme. Here we show that, in the reactions with aromatic and dicarboxylic ligands, the pK(a) of the Schiff base formed between the coenzyme pyridoxal 5'-phosphate and Lys258 or the substrate increases successively from 6.6 in the unliganded enzyme to approximately 8.8 in the Michaelis complex and to >10.5 in the external Schiff base complex. Mutations of Arg292 and Arg386 to Leu, which mimic neutralization of the positive charges of the two arginine residues by the ligand carboxylate groups, increased the Schiff base pK(a) by 0.1 and 0.7 unit, respectively. In contrast to these moderate effects of the Arg mutations, the cleavage of the Lys258 side chain of the Schiff base, which was brought about by preparing a mutant enzyme in which Lys258 was changed to Ala and the Schiff base was reconstituted with methylamine, produced the Schiff base pK(a) value of 10.2, that being 3.6 units higher than that of the wild-type enzyme. The observation indicates that the Schiff base pK(a) in the enzyme is lowered by the torsion around the C4-C4' axis of the Schiff base and suggests that the pK(a) is mainly controlled by changing the torsion angle during the course of catalysis. This mechanism, first observed for the reaction of aspartate aminotransferase with aspartate [Hayashi, H., Mizuguchi, H., and Kagamiyama, H. (1998) Biochemistry 37, 15076-15085], does not require the electrostatic contribution from the omega-carboxylate group of the substrate, and can explain why in aromatic amino acid aminotransferase the aromatic substrates can increase the Schiff base pK(a) during catalysis to the same extent as the dicarboxylic substrates. This is the first example in which the torsion pK(a) coupling of the pyridoxal 5'-phosphate Schiff base has been demonstrated in pyridoxal enzymes other than aspartate aminotransferase, and suggests the generality of the mechanism in the catalysis of aminotransferases related to aspartate aminotransferase.  相似文献   

11.
gamma-Aminobutyrate aminotransferase (GABA-AT), a pyridoxal phosphate-dependent enzyme, is responsible for the degradation of the inhibitory neurotransmitter GABA and is a target for antiepileptic drugs because its selective inhibition raises GABA concentrations in brain. The X-ray structure of pig GABA-AT has been determined to 3.0 A resolution by molecular replacement with the distantly related enzyme ornithine aminotransferase. Both omega-aminotransferases have the same fold, but exhibit side chain replacements in the closely packed binding site that explain their respective specificities. The aldimines of GABA and the antiepileptic drug vinyl-GABA have been modeled into the active site.  相似文献   

12.
The active site of microsomal aminopeptidase has been probed by studying the inhibition of the enzyme in the simultaneous presence of two ligands. The results have been analyzed with the Yonetani-Theorell plot to quantitate the degree of interaction between the two inhibitors. As expected, the enzyme contains a strong binding site for the alpha-amino group and the hydrophobic side chain of specific substrates. In addition, however, the enzyme can interact with another amine and a second hydrophobic group. Evidence suggests that this extra amine may bind to the zinc in an unprotonated form and that one of the hydrophobic sites is located in the vicinity. Another unexpected finding in this work is a strong synergism between the binding of ammonia and that of zinc ligands such as hydroxamates. This synergism may reflect an induced-fit mechanism that brings the catalytically important zinc atom into the optimal state only in the presence of specific substrates.  相似文献   

13.
A multiple sequence alignment among aspartate aminotransferase, dialkylglycine decarboxylase, and serine hydroxymethyltransferase (DAS) was used for profile databank search. The DAS profile could detect similarities to other pyridoxal or pyridoxamine phosphate-dependent enzymes, like several gene products involved in dideoxysugar and deoxyaminosugar synthesis. The alignment among DAS and such gene products shows the conservation of aspartate 222 and lysine 258, which, in aspartate aminotransferase, interacts with the N1 of the coenzyme pyridine ring and forms the internal Schiff base, respectively. The lysine is replaced by histidine in the pyridoxamine phosphate-dependent gene products. The alignment indicates also that the region encompassing the coenzyme binding site is the most conserved.  相似文献   

14.
Zhou X  Jin X  Medhekar R  Chen X  Dieckmann T  Toney MD 《Biochemistry》2001,40(5):1367-1377
The two half-reactions of the pyridoxal 5'-phosphate (PLP)-dependent enzyme dialkylglycine decarboxylase (DGD) were studied individually by multiwavelength stopped-flow spectroscopy. Biphasic behavior was found for the reactions of DGD-PLP, consistent with two coexisting conformations observed in steady-state kinetics [Zhou, X., and Toney, M. D. (1998) Biochemistry 37, 5761--5769]. The half-reaction kinetic parameters depend on alkali metal ion size in a manner similar to that observed for steady-state kinetic parameters. The fast phase maximal rate constant for the 2-aminoisobutyrate (AIB) decarboxylation half-reaction with the potassium form of DGD-PLP is 25 s(-1), while that for the transamination half-reaction between DGD-PMP and pyruvate is 75 s(-1). The maximal rate constant for the transamination half-reaction of the potassium form of DGD-PLP with L-alanine is 24 s(-1). The spectral data indicate that external aldimine formation with either AIB or L-alanine and DGD-PLP is a rapid equilibrium process, as is ketimine formation from DGD-PMP and pyruvate. Absorption ascribable to the quinonoid intermediate is not observed in the AIB decarboxylation half-reaction, but is observed in the dead-time of the stopped-flow in the L-alanine transamination half-reaction. The [1-(13)C]AIB kinetic isotope effect (KIE) on k(cat) for the steady-state reaction is 1.043 +/- 0.003, while a value of 1.042 +/- 0.009 was measured for the AIB half-reaction. The secondary KIE measured for the AIB decarboxylation half-reaction with [C4'-(2)H]PLP is 0.92 +/- 0.02. The primary [2-(2)H]-L-alanine KIE on the transamination half-reaction is unity. Small but significant solvent KIEs are observed on k(cat) and k(cat)/K(M) for both substrates, and the proton inventories are linear in each case. NMR measurements of C2--H washout vs product formation give ratios of 105 and 14 with L-alanine and isopropylamine as substrates, respectively. These results support a rate-limiting, concerted C alpha-decarboxylation/C4'-protonation mechanism for the AIB decarboxylation reaction, and rapid equilibrium quinonoid formation followed by rate-limiting protonation to the ketimine intermediate for the L-alanine transamination half-reaction. Energy profiles for the two half-reactions are constructed.  相似文献   

15.
A novel method is described for the cyclization of peptides--or segments of polypeptides--which requires a free N-terminal alpha-amino group and a distal amino acid residue containing a nucleophilic side chain. The reaction is conducted in two steps, both in the aqueous phase. The first step involves acylation of the N-terminal alpha-amino group with iodoacetic anhydride at pH 6. This acylation reaction has greater than 90% specificity for peptide alpha-amino groups and gives no alkylation of Arg, His, Lys or Met by the iodoacetate side product (R. Wetzel et al., Bioconjugate Chem., 1, 114-122, 1990). In the second step, the acylation reaction mixture or the isolated iodoacetyl-peptide is incubated at room temperature to give the cyclic peptide formed by reaction of the nucleophilic side chain with the iodoacetyl moiety. The pH dependence of the cyclization reaction by Met, Lys, Arg or His is consistent with the pKa of the nucleophilic side chain. Thus, peptides containing Met plus other nucleophilic amino acids should preferentially cyclize via Met at low pH. In this paper, preparation of cyclic peptides containing 3-6 amino acids is described; the full range of ring sizes and sequences which can undergo this cyclization has not been further explored. Preliminary results suggest that this method is also fairly general with respect to the amino acid sequence being cyclized. The reaction appears to be particularly suited for cyclization via Lys and Met side chains. All of the cyclized products are sufficiently stable for many biological applications.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

16.
A transport double mutant of Neurospora crassa has been isolated that has only one of the three transport systems capable of l-histidine uptake. The substrate specificity of the remaining transport system, termed the general transport system, has been fully characterized with regard to the contributions to binding of the side chain, the alpha-amino group, and the carboxylate group. The positively charged alpha-amino group is necessary for binding; the negatively charged carboxylate group is of less importance, since its replacement by a neutral carbonyl functional group does not completely abolish binding. The greatest structural latitude for binding was found in the side chain; affinity for alpha-amino acids was uniformly high except for l-aspartic and l-glutamic acids, l-asparagine, and l-proline. Thus, this transport system is "general" with these restrictions.  相似文献   

17.
Cytochrome P450 isolated from Bacillus subtilis (P450(BSbeta); molecular mass, 48 kDa) catalyzes the hydroxylation of a long-chain fatty acid (e.g. myristic acid) at the alpha- and beta-positions using hydrogen peroxide as an oxidant. We report here on the crystal structure of ferric P450(BSbeta) in the substrate-bound form, determined at a resolution of 2.1 A. P450(BSbeta) exhibits a typical P450 fold. The substrate binds to a specific channel in the enzyme and is stabilized through hydrophobic interactions of its alkyl side chain with some hydrophobic residues on the enzyme as well as by electrostatic interaction of its terminal carboxylate with the Arg(242) guanidium group. These interactions are responsible for the site specificity of the hydroxylation site in which the alpha- and beta-positions of the fatty acid come into close proximity to the heme iron sixth site. The fatty acid carboxylate group interacts with Arg(242) in the same fashion as has been reported for the active site of chloroperoxidase, His(105)-Glu(183), which is an acid-base catalyst in the peroxidation reactions. On the basis of these observations, a possible mechanism for the hydroxylation reaction catalyzed by P450(BSbeta) is proposed in which the carboxylate of the bound-substrate fatty acid assists in the cleavage of the peroxide O-O bond.  相似文献   

18.
Onesti S  Desogus G  Brevet A  Chen J  Plateau P  Blanquet S  Brick P 《Biochemistry》2000,39(42):12853-12861
Lysyl-tRNA synthetase is a member of the class II aminoacyl-tRNA synthetases and catalyses the specific aminoacylation of tRNA(Lys). The crystal structure of the constitutive lysyl-tRNA synthetase (LysS) from Escherichia coli has been determined to 2.7 A resolution in the unliganded form and in a complex with the lysine substrate. A comparison between the unliganded and lysine-bound structures reveals major conformational changes upon lysine binding. The lysine substrate is involved in a network of hydrogen bonds. Two of these interactions, one between the alpha-amino group and the carbonyl oxygen of Gly 216 and the other between the carboxylate group and the side chain of Arg 262, trigger a subtle and complicated reorganization of the active site, involving the ordering of two loops (residues 215-217 and 444-455), a change in conformation of residues 393-409, and a rotation of a 4-helix bundle domain (located between motif 2 and 3) by 10 degrees. The result of these changes is a closing up of the active site upon lysine binding.  相似文献   

19.
The X-ray crystal structure of the human alpha-thrombin-hirunorm IV complex has been determined at 2.5 A resolution, and refined to an R-factor of 0.173. The structure reveals an inhibitor binding mode distinctive of a true hirudin mimetic, which justifies the high inhibitory potency and the selectivity of hirunorm IV. This novel inhibitor, composed of 26 amino acids, interacts through the N-terminal end with the alpha-thrombin active site in a nonsubstrate mode, and binds specifically to the fibrinogen recognition exosite through the C-terminal end. The backbone of the N-terminal tripeptide Chg1"-Arg2"-2Na13" (Chg, cyclohexyl-glycine; 2Na1, beta-(2-naphthyl)-alanine) forms a parallel beta-strand to the thrombin main-chain segment Ser214-Gly216. The Chg1" side chain occupies the S2 site, Arg2" penetrates into the S1 specificity site, while the 2Na13" side chain occupies the aryl binding site. The Arg2" side chain enters the S1 specificity pocket from a position quite apart from the canonical P1 site. This notwithstanding, the Arg2" side chain establishes the typical ion pair with the carboxylate group of Asp189.  相似文献   

20.
Plasma amine oxidases catalyze the oxidative deamination of amines to aldehydes, followed by a 2e- reduction of O2 to H2O2. Pyrroloquinoline quinone (PQQ), previously believed to be restricted to prokaryotes, has recently been proposed to be the cofactor undergoing reduction in the first half-reaction of bovine plasma amine oxidase (Ameyama, M., Hayashi, U., Matsushita, K., Shinagawa, E., and Adachi, O. (1984) Agric. Biol. Chem. 48, 561-565; Lobenstein-Verbeek, C. L., Jongejan, J. A., Frank, J., and Duine, J. A. (1984) FEBS Lett. 170, 305-309). This result is unexpected, since model studies with PQQ implicate Schiff's base formation between a reactive carbonyl and substrates, whereas experiments with bovine plasma amine oxidase have failed to provide evidence for a carbonyl cofactor. We have, therefore, re-examined putative adducts between substrate and enzyme-bound cofactor, employing a combination of [14C]benzylamine and [3H]NaCNBH3. The use of the relatively weak reductant, NaCNBH3, affords Schiff's base specificity and permits the study of enzyme below pH 7.0. As we show, enzyme can only be inactivated by NaCNBH3 in the presence of substrate, leading to the incorporation of 1 mol of [14C]benzylamine/mol of enzyme subunit at complete inactivation. By contrast, we are unable to detect any labeling with [3H]NaCNBH3, analogous to an earlier study with [3H]NaCNBH4 (Suva, R. H., and Abeles, R. H. (1978) Biochemistry 17, 3538-3545). We conclude, first, that our inability to obtain adducts containing both carbon 14 and tritium rules out the reductive trapping either of amine substrate with pyridoxal phosphate or of aldehyde product with a lysyl side chain and, second, that the observed pattern of labeling is fully consistent with the presence of PQQ at the active site of bovine plasma amine oxidase.  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