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.
Copper amine oxidases are homodimeric enzymes that catalyze two reactions: first, a self-processing reaction to generate the 2,4,5-trihydroxyphenylalanine (TPQ) cofactor from an active site tyrosine by a single turnover mechanism; second, the oxidative deamination of primary amine substrates with the production of aldehyde, hydrogen peroxide, and ammonia catalyzed by the mature enzyme. The importance of active site residues in both of these processes has been investigated by structural studies and site-directed mutagenesis in enzymes from various organisms. One conserved residue is a tyrosine, Tyr369 in the Escherichia coli enzyme, whose hydroxyl is hydrogen bonded to the O4 of TPQ. To explore the importance of this site, we have studied a mutant enzyme in which Tyr369 has been mutated to a phenylalanine. We have determined the X-ray crystal structure of this variant enzyme to 2.1 A resolution, which reveals that TPQ adopts a predominant nonproductive conformation in the resting enzyme. Reaction of the enzyme with the irreversible inhibitor 2-hydrazinopyridine (2-HP) reveals differences in the reactivity of Y369F compared with wild type with more efficient formation of an adduct (lambda(max) = 525 nm) perhaps reflecting increased mobility of the TPQ adduct within the active site of Y369F. Titration with 2-HP also reveals that both wild type and Y369F contain one TPQ per monomer, indicating that Tyr369 is not essential for TPQ formation, although we have not measured the rate of TPQ biogenesis. The UV-vis spectrum of the Y369F protein shows a broader peak and red-shifted lambda(max) at 496 nm compared with wild type (480 nm), consistent with an altered electronic structure of TPQ. Steady-state kinetic measurements reveal that Y369F has decreased catalytic activity particularly below pH 6.5 while the K(M) for substrate beta-phenethylamine increases significantly, apparently due to an elevated pK(a) (5.75-6.5) for the catalytic base, Asp383, that should be deprotonated for efficient binding of protonated substrate. At pH 7.0, the K(M) for wild type and Y369F are similar at 1.2 and 1.5 microM, respectively, while k(cat) is decreased from 15 s(-1) in wild type to 0.38 s(-1), resulting in a 50-fold decrease in k(cat)/K(M) for Y369F. Transient kinetics experiments indicate that while the initial stages of enzyme reduction are slower in the variant, these do not represent the rate-limiting step. Previous structural and solution studies have implicated Tyr369 as a component of a proton shuttle from TPQ to dioxygen. The moderate changes in kinetic parameters observed for the Y369F variant indicate that if this is the case, then the absence of the Tyr369 hydroxyl can be compensated for efficiently within the active site.  相似文献   

2.
2-Hydrazinopyridine (2HP) is an irreversible inhibitor of copper amine oxidases (CAOs). 2HP reacts directly at the C5 position of the TPQ cofactor, yielding an intense chromophore with lambda(max) approximately 430 nm (adduct I) in Escherichia coli amine oxidase (ECAO). The adduct I form of wild type (WT-ECAO) was assigned as a hydrazone on the basis of the X-ray crystal structure. The hydrazone adduct appears to be stabilized by two key hydrogen-bonding interactions between the TPQ-2HP moiety and two active site residues: the catalytic base (D383) and the conserved tyrosine residue (Y369). In this work, we have synthesized a model compound (2) for adduct I from the reaction of a TPQ model compound (1) and 2HP. NMR spectroscopy and X-ray crystallography show that 2 exists predominantly as the azo form (lambda(max) at 414 nm). Comparison of the UV-vis and resonance Raman spectra of 2 with adduct I in WT, D383E, D383N, and Y369F forms of ECAO revealed that adduct I in WT and D383N is a tautomeric mixture where the hydrazone form is favored. In D383E adduct I, the equilibrium is further shifted in favor of the hydrazone form. UV-vis spectroscopic pH titrations of adduct I in WT, D383N, D383E, and 2 confirmed that D383 in WT adduct I is protonated at pH 7 and stabilizes the hydrazone tautomer by a short hydrogen-bonding interaction. The deprotonation of D383 (pKa approximately 9.7) in adduct I resulted in conversion of adduct I to the azo tautomer with a blue shift of the lambda(max) to 420 nm, close to that of 2. In contrast, adduct I in D383N and D383E is stable and did not show any pH-dependent spectral changes. In Y369F, adduct I was not stable and gradually converted into a new species with lambda(max) at approximately 530 nm (adduct II). A detailed mechanism for the adduct I formation in WT has been proposed that is consistent with the mechanism proposed for the oxidation of substrate by CAOs but addresses some key differences in the active site chemistry of hydrazine inhibitors and substrate amines.  相似文献   

3.
Copper amine oxidase contains a post-translationally generated quinone cofactor, topa quinone (TPQ), which mediates electron transfer from the amine substrate to molecular oxygen. The overall catalytic reaction is divided into the former reductive and the latter oxidative half-reactions based on the redox state of TPQ. In the reductive half-reaction, substrate amine reacts with the C5 carbonyl group of the oxidized TPQ, forming the substrate Schiff base (TPQ(ssb)), which is then converted to the product Schiff base (TPQ(psb)). During this step, an invariant Asp residue with an elevated pKa is presumed to serve as a general base accepting the alpha proton of the substrate. When Asp298, the putative active-site base in the recombinant enzyme from Arthrobacter globiformis, was mutated into Ala, the catalytic efficiency dropped to a level of about 10(6) orders of magnitude smaller than the wild-type (WT) enzyme, consistent with the essentiality of Asp298. Global analysis of the slow UV/vis spectral changes observed during the reductive half-reaction of the D298A mutant with 2-phenylethylamine provided apparent rate constants for the formation and decay of TPQ(ssb) (k(obs) = 4.7 and 4.8 x 10(-4) s(-1), respectively), both of which are markedly smaller than those of the WT enzyme determined by rapid-scan stopped-flow analysis (k(obs) = 699 and 411 s(-1), respectively). Thus, Asp298 plays important roles not only in the alpha-proton abstraction from TPQ(ssb) but also in other steps in the reductive half-reaction. X-ray diffraction analyses of D298A crystals soaked with the substrate for 1 h and 1 week revealed the structures of TPQ(ssb) and TPQ(psb), respectively, as pre-assigned by single-crystal microspectrophotometry. Consistent with the stereospecificity of alpha-proton abstraction, the pro-S alpha-proton of TPQ(ssb) to be abstracted is positioned nearly perpendicularly to the plane formed by the Schiff-base imine double bond conjugating with the quinone ring of TPQ, so that the orbitals of sigma and pi electrons maximally overlap in the conjugate system. More intriguingly, the pro-S alpha proton of the substrate is released stereospecifically even in the reaction catalyzed by the base-lacking D298A mutant. On the basis of these results, we propose that the stereospecificity of alpha-proton abstraction is primarily determined by the conformation of TPQ(ssb), rather than the relative geometry of TPQ and the catalytic base.  相似文献   

4.
Lie MA  Celik L  Jørgensen KA  Schiøtt B 《Biochemistry》2005,44(45):14792-14806
We have performed long-term molecular dynamics simulations of pyruvate decarboxylase from Zymomonas mobilis. Nine structures were modeled to investigate mechanistic questions related to binding of the cofactor, thiamin diphosphate (ThDP), and the substrate in the active site. The simulations reveal that the proposed three ThDP-tautomers all can bind in the active site and indicate that the equilibrium is shifted toward 4'-aminopyrimidine ThDP in the absence of substrate. 4'-Aminopyrimidinium ThDP is found to be a likely intermediate in the equilibrium. Mutations of important active site residues, Glu473Ala and Glu50Ala, were modeled to further elucidate their catalytic role. Formation of the catalytic important ylide by deprotonation of ThDP(C2) is investigated. Only the less favored tautomer, 1',4'-iminopyrimidine ThDP (imino-ThDP), could be deprotonated. The two other tautomers of ThDP could not be activated at the C2-position, thus, explaining the mechanistic importance of the less stable imino-ThDP. Finally, binding of pyruvate in the active site with the cofactor modeled as the nucleophilic ylide (ylide-ThDP) is studied. The carbonyl group of the substrate forms a hydrogen bond to Tyr290(OH). No hydrogen bond could be identified between ThDP(N4') and the substrate. The geometry of the substrate binding is well-suited for a nucleophilic attack by ylide-ThDP(C2). We propose that a proton relay from His113 via Asp27 and Tyr290 to the carbonyl oxygen atom of the substrate may be involved in the mechanism.  相似文献   

5.
Copper amine oxidases (CuAO), from Escherichia coli (ECAO) and pea seedling (PSAO) were reacted with an excess of the hydrazine derivative 2-hydrazinopyridine (2HP) to form an initial, strongly absorbing adduct, (adduct 1; λmax 420–430 nm) formed by the covalent binding of 2HP with the active site cofactor 2,4,5-trihydroxyphenylalanine quinone (TPQ). Thermal incubation of buffered solutions of adduct 1 (pH 5.65–10.7) or addition of KOH solution (giving a final pH of 13–15) led isosbestically to a dramatic λmax shift yielding adduct 2 (λmax 520–530 nm). For both ECAO and PSAO, an increase in pH resulted in increased formation of adduct 2 with concomitant loss of adduct 1. Maximum adduct 2 formation occurred at pH 9.84 in ECAO and at pH 10.7 in PSAO. Beyond these pH levels, adduct 2 formation occurred to a much lesser extent which was independent of pH, suggesting enzyme denaturation. It is proposed that the conversion of adduct 1 to adduct 2 occurs as a result of hydrazone to azo conversion mediated by loss of a single proton, possibly to the active site base. It is further postulated that adduct formation and subsequent deprotonation can be likened to the substrate and product Schiff base complexes in the reductive half cycle of copper/TPQ containing amine oxidases. As part of this study an extinction coefficient at 280 nm was determined for ECAO by gravimetric analysis. This yielded a value of 2.1×105 M−1 cm−1 giving rise to the need of a correction factor when estimating the protein concentration from an absorbance reading at 280 nm. Using the estimated molecular mass of 160 kDa for the homodimeric ECAO, a correction factor of 0.76 must be applied.  相似文献   

6.
Adduct I (lambda(max) at approximately 430 nm) formed in the reaction of 2-hydrazinopyridine (2HP) and the TPQ cofactor of wild-type Escherichia coli copper amine oxidase (WT-ECAO) is stable at neutral pH, 25 degrees C, but slowly converts to another spectroscopically distinct species with a lambda(max) at approximately 530 nm (adduct II) at pH 9.1. The conversion was accelerated either by incubation of the reaction mixture at 60 degrees C or by increasing the pH (>13). The active site base mutant forms of ECAO (D383N and D383E) showed spectral changes similar to WT when incubated at 60 degrees C. By contrast, in the Y369F mutant adduct I was not stable at pH 7, 25 degrees C, and gradually converted to adduct II, and this rate of conversion was faster at pH 9. To identify the nature of adduct II, we have studied the effects of pH and divalent cations on the UV-vis and resonance Raman spectroscopic properties of the model compound of adduct I (2). Strikingly, it was found that addition of Cu2+ to 2 at pH 7 gave a product (3) that exhibited almost identical spectroscopic signatures to adduct II. The X-ray crystal structure of 3 shows that it is the copper-coordinated form of 2, where the +2 charge of copper is neutralized by a double deprotonation of 2. These results led to the proposal that adduct II in the enzyme is TPQ-2HP that has migrated onto the active site Cu2+. The X-ray crystal structure of Y369F adduct II confirmed this assignment. Resonance Raman and EPR spectroscopy showed that adduct II in WT-ECAO is identical to that seen in Y369F. This study clearly demonstrates that the hydrogen-bonding interaction between O4 of TPQ and the conserved Tyr (Y369) is important in controlling the position and orientation of TPQ in the catalytic cycle, including optimal orientation for reactivity with substrate amines.  相似文献   

7.
The structure of amylosucrase from Neisseria polysaccharea in complex with beta-D-glucose has been determined by X-ray crystallography at a resolution of 1.66 A. Additionally, the structure of the inactive active site mutant Glu328Gln in complex with sucrose has been determined to a resolution of 2.0 A. The D-glucose complex shows two well-defined D-glucose molecules, one that binds very strongly in the bottom of a pocket that contains the proposed catalytic residues (at the subsite -1), in a nonstrained (4)C(1) conformation, and one that binds in the packing interface to a symmetry-related molecule. A third weaker D-glucose-binding site is located at the surface near the active site pocket entrance. The orientation of the D-glucose in the active site emphasizes the Glu328 role as the general acid/base. The binary sucrose complex shows one molecule bound in the active site, where the glucosyl moiety is located at the alpha-amylase -1 position and the fructosyl ring occupies subsite +1. Sucrose effectively blocks the only visible access channel to the active site. From analysis of the complex it appears that sucrose binding is primarily obtained through enzyme interactions with the glucosyl ring and that an important part of the enzyme function is a precise alignment of a lone pair of the linking O1 oxygen for hydrogen bond interaction with Glu328. The sucrose specificity appears to be determined primarily by residues Asp144, Asp394, Arg446, and Arg509. Both Asp394 and Arg446 are located in an insert connecting beta-strand 7 and alpha-helix 7 that is much longer in amylosucrase compared to other enzymes from the alpha-amylase family (family 13 of the glycoside hydrolases).  相似文献   

8.
Quinohemoprotein amine dehydrogenase (QH-AmDH) catalyzes the oxidative deamination of aliphatic and aromatic amines. The enzyme from Pseudomonas putida has an alpha beta gamma heterotrimeric structure with two heme c groups in the largest alpha subunit, and a novel quinone cofactor [cysteine tryptophylquinone (CTQ)] and hitherto unknown internal cross-bridges in the smallest gamma subunit. The crystal structure of the enzyme in the complex with the inhibitor [p-nitrophenylhydrazine (pNPH)] has been determined at a 2.0 A resolution.(1) The hydrazone of the cofactor with the inhibitor was nicely modeled into the omit electron density map, identifying the C6 carbonyl group as the reactive site of the cofactor. The Asp33 gamma is unambiguously determined as the catalytic base to abstract the alpha-proton from a substrate, because N beta atom of the inhibitor corresponding to the C alpha atom of the substrate amine is neighbored to Asp33 gamma. The bound inhibitor is completely enclosed in the active site pocket formed by the residues from the beta- and gamma-subunits. The cofactor-inhibitor adduct may be predominantly in the hydrazone with the azo form as a minor component. The binding of the inhibitor causes minor but important conformational changes in the residues surrounding the active site. The inhibitor may have access to the active site pocket through the water-filled crevice between the beta- and gamma-subunits.  相似文献   

9.
Bacterial RibG is a potent target for antimicrobial agents, because it catalyzes consecutive deamination and reduction steps in the riboflavin biosynthesis. In the N-terminal deaminase domain of Bacillus subtilis RibG, structure-based mutational analyses demonstrated that Glu51 and Lys79 are essential for the deaminase activity. In the C-terminal reductase domain, the complex structure with the substrate at 2.56-A resolution unexpectedly showed a ribitylimino intermediate bound at the active site, and hence suggested that the ribosyl reduction occurs through a Schiff base pathway. Lys151 seems to have evolved to ensure specific recognition of the deaminase product rather than the substrate. Glu290, instead of the previously proposed Asp199, would seem to assist in the proton transfer in the reduction reaction. A detailed comparison reveals that the reductase and the pharmaceutically important enzyme, dihydrofolate reductase involved in the riboflavin and folate biosyntheses, share strong conservation of the core structure, cofactor binding, catalytic mechanism, even the substrate binding architecture.  相似文献   

10.
Copper amine oxidases (CAOs) catalyse the oxidation of various aliphatic amines to the corresponding aldehydes, ammonia and hydrogen peroxide. Although CAOs from various organisms share a highly conserved active-site structure including a protein-derived cofactor, topa quinone (TPQ), their substrate specificities differ considerably. To obtain structural insights into the substrate specificity of a CAO from Arthrobacter globiformis (AGAO), we have determined the X-ray crystal structures of AGAO complexed with irreversible inhibitors that form covalent adducts with TPQ. Three hydrazine derivatives, benzylhydrazine (BHZ), 4-hydroxybenzylhydrazine (4-OH-BHZ) and phenylhydrazine (PHZ) formed predominantly a hydrazone adduct, which is structurally analogous to the substrate Schiff base of TPQ formed during the catalytic reaction. With BHZ and 4-OH-BHZ, but not with PHZ, the inhibitor aromatic ring is bound to a hydrophobic cavity near the active site in a well-defined conformation. Furthermore, the hydrogen atom on the hydrazone nitrogen is located closer to the catalytic base in the BHZ and 4-OH-BHZ adducts than in the PHZ adduct. These results correlate well with the reactivity of 2-phenylethylamine and tyramine as preferred substrates for AGAO and also explain why benzylamine is a poor substrate with markedly decreased rate constants for the steps of proton abstraction and the following hydrolysis.  相似文献   

11.
The enzyme 6-pyruvoyl tetrahydropterin synthase (PTPS) catalyses the second step in the de novo biosynthesis of tetrahydrobiopterin, the conversion of dihydroneopterin triphosphate to 6-pyruvoyl tetrahydropterin. The Zn and Mg-dependent reaction includes a triphosphate elimination, a stereospecific reduction of the N5-C6 double bond and the oxidation of both side-chain hydroxyl groups. The crystal structure of the inactive mutant Cys42Ala of PTPS in complex with its natural substrate dihydroneopterinetriphosphate was determined at 1.9 A resolution. Additionally, the uncomplexed enzyme was refined to 2.0 A resolution. The active site of PTPS consists of the pterin-anchoring Glu A107 neighboured by two catalytic motifs: a Zn(II) binding site and an intersubunit catalytic triad formed by Cys A42, Asp B88 and His B89. In the free enzyme the Zn(II) is in tetravalent co-ordination with three histidine ligands and a water molecule. In the complex the water is replaced by the two substrate side-chain hydroxyl groups yielding a penta-co-ordinated Zn(II) ion. The Zn(II) ion plays a crucial role in catalysis. It activates the protons of the substrate, stabilizes the intermediates and disfavours the breaking of the C1'C2' bond in the pyruvoyl side-chain. Cys A42 is activated by His B89 and Asp B88 for proton abstraction from the two different substrate side-chain atoms C1', and C2'. Replacing Ala A42 in the mutant structure by the wild-type Cys by modelling shows that the C1' and C2' substrate side-chain protons are at equal distances to Cys A42 Sgamma. The basicity of Cys A42 may be increased by a catalytic triad His B89 and Asp B88. The active site of PTPS seems to be optimised to carry out proton abstractions from two different side-chain C1' and C2' atoms, with no obvious preference for one of them. Kinetic studies with dihydroneopterin monophosphate reveal that the triphosphate moiety of the substrate is necessary for enzyme specifity.  相似文献   

12.
Hegeman AD  Gross JW  Frey PA 《Biochemistry》2001,40(22):6598-6610
A model of the Escherichia coli dTDP-glucose-4,6-dehydratase (4,6-dehydratase) active site has been generated by combining amino acid sequence alignment information with the 3-dimensional structure of UDP-galactose-4-epimerase. The active site configuration is consistent with the partially refined 3-dimensional structure of 4,6-dehydratase, which lacks substrate-nucleotide but contains NAD(+) (PDB file ). From the model, two groups of active site residues were identified. The first group consists of Asp135(DEH), Glu136(DEH), Glu198(DEH), Lys199(DEH), and Tyr301(DEH). These residues are near the substrate-pyranose binding pocket in the model, they are completely conserved in 4,6-dehydratase, and they differ from the corresponding equally well-conserved residues in 4-epimerase. The second group of residues is Cys187(DEH), Asn190(DEH), and His232(DEH), which form a motif on the re face of the cofactor nicotinamide binding pocket that resembles the catalytic triad of cysteine-proteases. The importance of both groups of residues was tested by mutagenesis and steady-state kinetic analysis. In all but one case, a decrease in catalytic efficiency of approximately 2 orders of magnitude below wild-type activity was observed. Mutagenesis of each of these residues, with the exception of Cys187(DEH), which showed near-wild-type activity, clearly has important negative consequences for catalysis. The allocation of specific functions to these residues and the absolute magnitude of these effects are obscured by the complex chemistry in this multistep mechanism. Tools will be needed to characterize each chemical step individually in order to assign loss of catalytic efficiency to specific residue functions. To this end, the effects of each of these variants on the initial dehydrogenation step were evaluated using a the substrate analogue dTDP-xylose. Additional steady-state techniques were employed in an attempt to further limit the assignment of rate limitation. The results are discussed within the context of the 4,6-dehydratase active site model and chemical mechanism.  相似文献   

13.
The catalytic domain of chitobiase (beta-N-1-4 acetylhexosaminidase) from Serratia marcescens, is an alpha/beta TIM-barrel. This enzyme belongs to family 20 of glycosyl hydrolases in which a conserved amino acid pair, aspartate-glutamate, is present (Asp539-Glu540). It was proposed that catalysis by this enzyme family is carried out by glutamate 540 acting as a proton donor and by the acetamido group of the substrate as a nucleophile. We investigated the role of Asp539 and Glu540 by site-directed mutagenesis, biochemical characterization and by structural analyses of chitobiase -substrate co-crystals. We found that both residues are essential for chitobiase activity. The mutations, however, led to subtle changes in the catalytic site. Our results support the model that Glu540 acts as the proton donor and that Asp539 acts in several different ways. Asp539 restrains the acetamido group of the substrate in a specific orientation by forming a hydrogen bond with N2 of the non-reduced (-1) sugar. In addition, this residue participates in substrate binding. It is also required for the correct positioning of Glu540 and may provide additional negative charge at the active site. Thus, these biochemical and structural studies provide a molecular explanation for the functional importance and conservation of these residues.  相似文献   

14.
The thiamin diphosphate (ThDP)-dependent enzyme indolepyruvate decarboxylase (IPDC) is involved in the biosynthetic pathway of the phytohormone 3-indoleacetic acid and catalyzes the nonoxidative decarboxylation of 3-indolepyruvate to 3-indoleacetaldehyde and carbon dioxide. The steady-state distribution of covalent ThDP intermediates of IPDC reacting with 3-indolepyruvate and the alternative substrates benzoylformate and pyruvate has been analyzed by (1)H NMR spectroscopy. For the first time, we are able to isolate and directly assign covalent intermediates of ThDP with aromatic substrates. The intermediate analysis of IPDC variants is used to infer the involvement of active site side chains and functional groups of the cofactor in distinct catalytic steps during turnover of the different substrates. As a result, three residues (glutamate 468, aspartate 29, and histidine 115) positioned perpendicular to the thiazolium moiety of ThDP are involved in binding of all substrates and decarboxylation of the respective tetrahedral ThDP-substrate adducts. Most likely, interactions of these side chains with the substrate-derived carboxylate account for an optimal orientation of the substrate and/or intermediate in the course of carbon-carbon ligation and decarboxylation supporting the suggested least-motion, maximum overlap mechanism. The active site residue glutamine 383, which is located at the opposite site of the thiazolium nucleus as the "carboxylate pocket" (formed by the Glu-Asp-His triad), is central to the substrate specificity of IPDC, probably through orbital alignment. The Glu51-cofactor proton shuttle is, conjointly with the Glu-Asp-His triad, involved in multiple proton transfer steps, including ylide generation, substrate binding, and product release. Studies with para-substituted benzoylformate substrates demonstrate that the electronic properties of the substrate affect the stabilization or destabilization of the carbanion intermediate or carbanion-like transition state and in that way alter the rate dependence on decarboxylation. In conclusion, general mechanistic principles of catalysis of ThDP-dependent enzymes are discussed.  相似文献   

15.
The heme-containing catalase HPII of Escherichia coli consists of a homotetramer in which each subunit contains a core region with the highly conserved catalase tertiary structure, to which are appended N- and C-terminal extensions making it the largest known catalase. HPII does not bind NADPH, a cofactor often found in catalases. In HPII, residues 585-590 of the C-terminal extension protrude into the pocket corresponding to the NADPH binding site in the bovine liver catalase. Despite this difference, residues that define the NADPH pocket in the bovine enzyme appear to be well preserved in HPII. Only two residues that interact ionically with NADPH in the bovine enzyme (Asp212 and His304) differ in HPII (Glu270 and Glu362), but their mutation to the bovine sequence did not promote nucleotide binding. The active-site heme groups are deeply buried inside the molecular structure requiring the movement of substrate and products through long channels. One potential channel is about 30 A in length, approaches the heme active site laterally, and is structurally related to the branched channel associated with the NADPH binding pocket in catalases that bind the dinucleotide. In HPII, the upper branch of this channel is interrupted by the presence of Arg260 ionically bound to Glu270. When Arg260 is replaced by alanine, there is a threefold increase in the catalytic activity of the enzyme. Inhibitors of HPII, including azide, cyanide, various sulfhydryl reagents, and alkylhydroxylamine derivatives, are effective at lower concentration on the Ala260 mutant enzyme compared to the wild-type enzyme. The crystal structure of the Ala260 mutant variant of HPII, determined at 2.3 A resolution, revealed a number of local structural changes resulting in the opening of a second branch in the lateral channel, which appears to be used by inhibitors for access to the active site, either as an inlet channel for substrate or an exhaust channel for reaction products.  相似文献   

16.
The oxidation of IMP to XMP is the rate-limiting step in the de novo synthesis of guanine ribonucleotides. This NAD-dependent reaction is catalyzed by the enzyme inosine monophosphate dehydrogenase (IMPDH). Based upon the recent structural determination of IMPDH complexed to oxidized IMP (XMP*) and the potent uncompetitive inhibitor mycophenolic acid (MPA), we have selected active site residues and prepared mutants of human type II IMPDH. The catalytic parameters of these mutants were determined. Mutations G326A, D364A, and the active site nucleophile C331A all abolish enzyme activity to less than 0.1% of wild type. These residues line the IMP binding pocket and are necessary for correct positioning of the substrate, Asp364 serving to anchor the ribose ring of the nucleotide. In the MPA/NAD binding site, significant loss of activity was seen by mutation of any residue of the triad Arg322, Asn303, Asp274 which form a hydrogen bonding network lining one side of this pocket. From a model of NAD bound to the active site consistent with the mutational data, we propose that these resides are important in binding the ribose ring of the nicotinamide substrate. Additionally, mutations in the pair Thr333, Gln441, which lies close to the xanthine ring, cause a significant drop in the catalytic activity of IMPDH. It is proposed that these residues serve to deliver the catalytic water molecule required for hydrolysis of the cysteine-bound XMP* intermediate formed after oxidation by NAD.  相似文献   

17.
Glutamine synthetase (GS) catalyzes the ATP-dependent condensation of ammonia and glutamate to yield glutamine, ADP, and inorganic phosphate in the presence of divalent cations. Bacterial GS is an enzyme of 12 identical subunits, arranged in two rings of 6, with the active site between each pair of subunits in a ring. In earlier work, we have reported the locations within the funnel-shaped active site of the substrates glutamate and ATP and of the two divalent cations, but the site for ammonia (or ammonium) has remained elusive. Here we report the discovery by X-ray crystallography of a binding site on GS for monovalent cations, Tl+ and Cs+, which is probably the binding site for the substrate ammonium ion. Fourier difference maps show the following. (1) Tl+ and Cs+ bind at essentially the same site, with ligands being Glu 212, Tyr 179, Asp 50', Ser 53' of the adjacent subunit, and the substrate glutamate. From its position adjacent to the substrate glutamate and the cofactor ADP, we propose that this monovalent cation site is the substrate ammonium ion binding site. This proposal is supported by enzyme kinetics. Our kinetic measurements show that Tl+, Cs+, and NH4+ are competitive inhibitors to NH2OH in the gamma-glutamyl transfer reaction. (2) GS is a trimetallic enzyme containing two divalent cation sites (n1, n2) and one monovalent cation site per subunit. These three closely spaced ions are all at the active site: the distance between n1 and n2 is 6 A, between n1 and Tl+ is 4 A, and between n2 and Tl+ is 7 A. Glu 212 and the substrate glutamate are bridging ligands for the n1 ion and Tl+. (3) The presence of a monovalent cation in this site may enhance the structural stability of GS, because of its effect of balancing the negative charges of the substrate glutamate and its ligands and because of strengthening the "side-to-side" intersubunit interaction through the cation-protein bonding. (4) The presence of the cofactor ADP increases the Tl+ binding to GS because ADP binding induces movement of Asp 50' toward this monovalent cation site, essentially forming the site. This observation supports a two-step mechanism with ordered substrate binding: ATP first binds to GS, then Glu binds and attacks ATP to form gamma-glutamyl phosphate and ADP, which complete the ammonium binding site. The third substrate, an ammonium ion, then binds to GS, and then loses a proton to form the more active species ammonia, which attacks the gamma-glutamyl phosphate to yield Gln. (5) Because the products (Glu or Gln) of the reactions catalyzed by GS are determined by the molecule (water or ammonium) attacking the intermediate gamma-glutamyl phosphate, this negatively charged ammonium binding pocket has been designed naturally for high affinity of ammonium to GS, permitting glutamine synthesis to proceed in aqueous solution.  相似文献   

18.
Site-directed mutagenesis has been used to explore the role of two carboxylates in the active site of histidine decarboxylase from Lactobacillus 30a. The most striking observation is that conversion of Glu197 to either Gln or Asp causes a major decrease in catalytic rate while enhancing substrate binding. This is consistent with models based on X-ray diffraction results which suggest that the acid may protonate a reaction intermediate during catalysis. The Asp197 protein undergoes a suicide reaction with substrate, apparently triggered by inappropriate protonation of the intermediate. This leads to decarboxylation-dependent transamination which converts the pyruvoyl cofactor to an alanine, inactivating the enzyme. Conversion of Glu66 to Gln affects parameters of kinetic cooperativity. The mutation fixes the Hill number at approximately 1.5, midway between the pH-dependent values of the wild-type enzyme.  相似文献   

19.
Shobe J  Dickinson CD  Ruf W 《Biochemistry》1999,38(9):2745-2751
Coagulation factor VIIa is an allosterically regulated trypsin-like serine protease that initiates the coagulation pathways upon complex formation with its cellular receptor and cofactor tissue factor (TF). The analysis of a conformation-sensitive monoclonal antibody directed to the macromolecular substrate exosite in the VIIa protease domain demonstrated a conformational link from this exosite to the catalytic cleft that is independent of cofactor-induced allosteric changes. In this study, we identify Glu 154 as a critical surface-exposed exosite residue side chain that undergoes conformational changes upon active site inhibitor binding. The Glu 154 side chain is important for hydrolysis of scissile bond mimicking peptidyl p-nitroanilide substrates, and for inhibition of VIIa's amidolytic function upon antibody binding. This exosite residue is not linked to the catalytic cleft residue Lys 192 which plays an important role in thrombin's allosteric coupling to exosite I. Allosteric linkages between VIIa's active site and the cofactor binding site or between the cofactor binding site and the macromolecular substrate exosite were not influenced by mutation of Glu 154. Glu 154 thus only influences the linkage of the macromolecular substrate binding exosite to the catalytic center. These data provide novel evidence that allosteric regulation of VIIa's catalytic function involves discrete and independent conformational linkages and that allosteric transitions in the VIIa protease domain are not globally coupled.  相似文献   

20.
Oxalate decarboxylase (EC 4.1.1.2) catalyzes the conversion of oxalate to formate and carbon dioxide and utilizes dioxygen as a cofactor. By contrast, the evolutionarily related oxalate oxidase (EC 1.2.3.4) converts oxalate and dioxygen to carbon dioxide and hydrogen peroxide. Divergent free radical catalytic mechanisms have been proposed for these enzymes that involve the requirement of an active site proton donor in the decarboxylase but not the oxidase reaction. The oxidase possesses only one domain and manganese binding site per subunit, while the decarboxylase has two domains and two manganese sites per subunit. A structure of the decarboxylase together with a limited mutagenesis study has recently been interpreted as evidence that the C-terminal domain manganese binding site (site 2) is the catalytic site and that Glu-333 is the crucial proton donor (Anand, R., Dorrestein, P. C., Kinsland, C., Begley, T. P., and Ealick, S. E. (2002) Biochemistry 41, 7659-7669). The N-terminal binding site (site 1) of this structure is solvent-exposed (open) and lacks a suitable proton donor for the decarboxylase reaction. We report a new structure of the decarboxylase that shows a loop containing a 3(10) helix near site 1 in an alternative conformation. This loop adopts a "closed" conformation forming a lid covering the entrance to site 1. This conformational change brings Glu-162 close to the manganese ion, making it a new candidate for the crucial proton donor. Site-directed mutagenesis of equivalent residues in each domain provides evidence that Glu-162 performs this vital role and that the N-terminal domain is either the sole or the dominant catalytically active domain.  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