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.
Tang KH  Harms A  Frey PA 《Biochemistry》2002,41(27):8767-8776
Lysine 5,6-aminomutase (5,6-LAM) catalyzes the interconversion of D-lysine with 2,5-diaminohexanoate and of L-beta-lysine with 3,5-diaminohexanoate. The coenzymes for 5,6-LAM are adenosylcobalamin (AdoCbl) and pyridoxal 5'-phosphate (PLP). In the proposed chemical mechanism, AdoCbl initiates the formation of substrate radicals, and PLP facilitates the radical rearrangement by forming an external aldimine linkage with the epsilon-amino group of a substrate, either D-lysine or L-beta-lysine. In the resting enzyme, an internal aldimine between PLP and an essential lysine in the active site facilitates productive PLP binding and catalysis. We present here biochemical, biophysical, and site-directed mutagenesis experiments, which document the existence of an essential lysine residue in the active site of 5,6-LAM from Porphyromonas gingivalis. Reduction of 5,6-LAM with NaBH(4) rapidly inactivates the enzyme and shifts the electronic absorption band from 420 to 325 nm. This is characteristic of the reduction of an aldimine linkage between the carbonyl group of PLP and the epsilon-amino group of a lysine residue. The reduced peptide was identified by Q-TOF/MS and further confirmed by Q-TOF/MS/MS sequencing. We show that lysine 144 in the small subunit of 5,6-LAM is the essential lysine residue. Lysine 144(beta) is separated by only 11 amino acids from histidine 133(beta), which forms a part of the "base-off"-AdoCbl binding motif. The sequence of the novel PLP-binding motif is conserved in 5,6-LAM from Clostridium sticklandii and P. gingivalis, and it is distinct from all known PLP-binding motifs. Mutation of lysine 144(beta) to glutamine led to K144Q(beta)-5,6-LAM, which displayed no enzymatic activity and no absorption band corresponding to an internal PLP-aldamine. In summary, we introduce a novel PLP-binding motif, the first to be discovered in an AdoCbl-dependent enzyme.  相似文献   

2.
Tang KH  Chang CH  Frey PA 《Biochemistry》2001,40(17):5190-5199
The lysine 5,6-aminomutase (5,6-LAM) purified from Clostridium sticklandii was found to undergo rapid inactivation in the absence of the activating enzyme E(2) and ATP. In the presence of substrate, inactivation was also seen for the recombinant 5,6-LAM. This adenosylcobalamin-dependent enzyme is postulated to generate cob(II)alamin and the 5'-deoxyadenosyl radical through enzyme-induced homolytic scission of the Co-C bond. However, the products cob(III)alamin and 5'-deoxyadenosine were observed upon inactivation of 5,6-LAM. Cob(III)alamin production, as monitored by the increase in A(358), proceeds at the same rate as the loss of enzyme activity, suggesting that the activity loss is related to the adventitious generation of cob(III)alamin during enzymatic turnover. The cleavage of adenosylcobalamin to cob(III)alamin is accompanied by the formation of 5'-deoxyadenosine at the same rate, and the generation of cob(III)alamin proceeds at the same rate both aerobically and anaerobically. Suicide inactivation requires the presence of substrate, adenosylcobalamin, and PLP. We have ruled out the involvement of either the putative 5'-deoxyadenosyl radical or dioxygen in suicide inactivation. We have shown that one or more reaction intermediates derived from the substrate or/and the product, presumably a radical, participate in suicide inactivation of 5,6-LAM through electron transfer from cob(II)alamin. Moreover, L-lysine is found to be a slowly reacting substrate, and it induces inactivation at a rate similar to that of D-lysine. The alternative substrate beta-lysine induces inactivation at least 25 times faster than DL-lysine. The inactivation mechanism is compatible with the radical isomerization mechanism proposed to explain the action of 5,6-LAM.  相似文献   

3.
The Staphylococcus aureus transpeptidase SrtA catalyzes the covalent attachment of LPXTG-containing virulence and colonization-associated proteins to cell-wall peptidoglycan in Gram-positive bacteria. Recent structural characterizations of staphylococcal SrtA, and related transpeptidases SrtB from S. aureus and Bacillus anthracis, provide many details regarding the active site environment, yet raise questions with regard to the nature of catalysis and active site cysteine thiol activation. Here we re-evaluate the kinetic mechanism of SrtA and shed light on aspects of its catalytic mechanism. Using steady-state, pre-steady-state, bisubstrate kinetic studies, and high-resolution electrospray mass spectrometry, revised steady-state kinetic parameters and a ping-pong hydrolytic shunt kinetic mechanism were determined for recombinant SrtA. The pH dependencies of kinetic parameters k(cat)/K(m) and k(cat) for the substrate Abz-LPETG-Dap(Dnp)-NH(2) were bell-shaped with pK(a) values of 6.3 +/- 0.2 and 9.4 +/- 0.2 for k(cat) and 6.2 +/- 0.2 and 9.4 +/- 0.2 for k(cat)/K(m). Solvent isotope effect (SIE) measurements revealed inverse behavior, with a (D)2(O)k(cat) of 0.89 +/- 0.01 and a (D)2(O)(k(cat)/K(m)) of 0.57 +/- 0.03 reflecting an equilibrium SIE. In addition, SIE measurements strongly implicated Cys184 participation in the isotope-sensitive rate-determining chemical step when considered in conjunction with an inverse linear proton inventory for k(cat). Last, the pH dependence of SrtA inactivation by iodoacetamide revealed a single ionization for inactivation. These studies collectively provide compelling evidence for a reverse protonation mechanism where a small fraction (ca. 0.06%) of SrtA is competent for catalysis at physiological pH, yet is highly active with an estimated k(cat)/K(m) of >10(5) M(-)(1) s(-)(1).  相似文献   

4.
Application of the amino acid analyzer is described for the separation, identification, and estimation of several uncommon dibasic amino acids that are eluted with a 0.35 m, pH 5.28, sodium citrate buffer between lysine and ammonia, namely, 2,5-diaminohexanoate, threo- and erythro-3,5-diaminohexanoate, 3,6-diaminohexanoate (β-lysine) and 2,4-diaminopentanoic acid. Conditions are also given for estimating d- and l-β-lysine as their l-glutamyl peptides, and for estimating the enantiomers of threo- and erythro-3,5-diaminohexanoate as the l-glutamyl derivative of the corresponding lactams.  相似文献   

5.
Catalytic reaction pathway for the mitogen-activated protein kinase ERK2   总被引:2,自引:0,他引:2  
Prowse CN  Hagopian JC  Cobb MH  Ahn NG  Lew J 《Biochemistry》2000,39(20):6258-6266
The structural, functional, and regulatory properties of the mitogen-activated protein kinases (MAP kinases) have long attracted considerable attention owing to the critical role that these enzymes play in signal transduction. While several MAP kinase X-ray crystal structures currently exist, there is by comparison little mechanistic information available to correlate the structural data with the known biochemical properties of these molecules. We have employed steady-state kinetic and solvent viscosometric techniques to characterize the catalytic reaction pathway of the MAP kinase ERK2 with respect to the phosphorylation of a protein substrate, myelin basic protein (MBP), and a synthetic peptide substrate, ERKtide. A minor viscosity effect on k(cat) with respect to the phosphorylation of MBP was observed (k(cat) = 10 +/- 2 s(-1), k(cat)(eta) = 0.18 +/- 0.05), indicating that substrate processing occurs via slow phosphoryl group transfer (12 +/- 4 s(-1)) followed by the faster release of products (56 +/- 4 s(-1)). At an MBP concentration extrapolated to infinity, no significant viscosity effect on k(cat)/K(m(ATP)) was observed (k(cat)/K(m(ATP)) = 0.2 +/- 0.1 microM(-1) s(-1), k(cat)/K(m(ATP))(eta) = -0.08 +/- 0.04), consistent with rapid-equilibrium binding of the nucleotide. In contrast, at saturating ATP, a full viscosity effect on k(cat)/K(m) for MBP was apparent (k(cat)/K(m(MBP)) = 2.4 +/- 1 microM(-1) s(-1), k(cat)/K(m(MBP))(eta) = 1.0 +/- 0.1), while no viscosity effect was observed on k(cat)/K(m) for the phosphorylation of ERKtide (k(cat)/K(m(ERKtide)) = (4 +/- 2) x 10(-3) microM(-1) s(-1), k(cat)/K(m(ERKtide))(eta) = -0.02 +/- 0.02). This is consistent with the diffusion-limited binding of MBP, in contrast to the rapid-equilibrium binding of ERKtide, to form the ternary Michaelis complex. Calculated values for binding constants show that the estimated value for K(d(MBP)) (/= 1.5 mM). The dramatically higher catalytic efficiency of MBP in comparison to that of ERKtide ( approximately 600-fold difference) is largely attributable to the slow dissociation rate of MBP (/=56 s(-1)), from the ERK2 active site.  相似文献   

6.
Histamine dehydrogenase (NSHADH) can be isolated from cultures of Nocardioides simplex grown with histamine as the sole nitrogen source. A previous report suggested that NSHADH might contain the quinone cofactor tryptophan tryptophyl quinone (TTQ). Here, the hdh gene encoding NSHADH is cloned from the genomic DNA of N. simplex, and the isolated enzyme is subjected to a full spectroscopic characterization. Protein sequence alignment shows NSHADH to be related to trimethylamine dehydrogenase (TMADH: EC 1.5.99.7), where the latter contains a bacterial ferredoxin-type [4Fe-4S] cluster and 6-S-cysteinyl FMN cofactor. NSHADH has no sequence similarity to any TTQ containing amine dehydrogenases. NSHADH contains 3.6+/-0.3 mol Fe and 3.7+/-0.2 mol acid labile S per subunit. A comparison of the UV/vis spectra of NSHADH and TMADH shows significant similarity. The EPR spectrum of histamine reduced NSHADH also supports the presence of the flavin and [4Fe-4S] cofactors. Importantly, we show that NSHADH has a narrow substrate specificity, oxidizing only histamine (K(m)=31+/-11 microM, k(cat)/K(m)=2.1 (+/-0.4)x10(5)M(-1)s(-1)), agmatine (K(m)=37+/-6 microM, k(cat)/K(m)=6.0 (+/-0.6)x10(4)M(-1)s(-1)), and putrescine (K(m)=1280+/-240 microM, k(cat)/K(m)=1500+/-200 M(-1)s(-1)). A kinetic characterization of the oxidative deamination of histamine by NSHADH is presented that includes the pH dependence of k(cat)/K(m) (histamine) and the measurement of a substrate deuterium isotope effect, (D)(k(cat)/K(m) (histamine))=7.0+/-1.8 at pH 8.5. k(cat) is also pH dependent and has a reduced substrate deuterium isotope of (D)(k(cat))=1.3+/-0.2.  相似文献   

7.
Rao S  Holz RC 《Biochemistry》2008,47(46):12057-12064
In order to gain insight into the catalytic mechanism of Fe-type nitrile hydratases (NHase), the pH and temperature dependence of the kinetic parameters k cat, K m, and k cat/ K m along with the solvent isotope effect were examined for the Fe-type NHase from Comamonas testosteroni Ni1 ( CtNHase). CtNHase was found to exhibit a bell-shaped curve for plots of relative activity vs pH over pH values 4-10 for the hydration of acrylonitrile and was found to display maximal activity at pH approximately 7.2. Fits of these data provided a p K ES1 value of 6.1 +/- 0.1, a p K ES2 value of 9.1 +/- 0.2 ( k' cat = 10.1 +/- 0.3 s (-1)), a p K E1 value of 6.2 +/- 0.1, and a p K E2 value of 9.2 +/- 0.1 ( k' cat/ K' m of 2.0 +/- 0.2 s (-1) mM (-1)). Proton inventory studies indicate that two protons are transferred in the rate-limiting step of the reaction at pH 7.2. Since CtNHase is stable to 25 degrees C, an Arrhenius plot was constructed by plotting ln( k cat) vs 1/ T, providing an E a of 33.3 +/- 1.5 kJ/mol. Delta H degrees of ionization values were also determined, thus helping to identify the ionizing groups exhibiting the p K ES1 and p K ES2 values. Based on Delta H degrees ion data, p K ES1 is assigned to betaTyr68 while p K ES2 is assigned to betaArg52, betaArg157, or alphaSer116 (NHases are alpha 2beta 2 heterotetramers). Given the strong similarities in the kinetic data obtained for both Co- and Fe-type NHase enzymes, both types of NHase enzymes likely hydrate nitriles in a similar fashion.  相似文献   

8.
Cathepsin C, or dipeptidyl peptidase I, is a lysosomal cysteine protease of the papain family that catalyzes the sequential removal of dipeptides from the free N-termini of proteins and peptides. Using the dipeptide substrate Ser-Tyr-AMC, cathepsin C was characterized in both steady-state and pre-steady-state kinetic modes. The pH(D) rate profiles for both log k cat/ K m and log k cat conformed to bell-shaped curves for which an inverse solvent kinetic isotope effect (sKIE) of 0.71 +/- 0.14 for (D)( k cat/ K a) and a normal sKIE of 2.76 +/- 0.03 for (D) k cat were obtained. Pre-steady-state kinetics exhibited a single-exponential burst of AMC formation in which the maximal acylation rate ( k ac = 397 +/- 5 s (-1)) was found to be nearly 30-fold greater than the rate-limiting deacylation rate ( k dac = 13.95 +/- 0.013 s (-1)) and turnover number ( k cat = 13.92 +/- 0.001 s (-1)). Analysis of pre-steady-state burst kinetics in D 2O allowed abstraction of a normal sKIE for the acylation half-reaction that was not observed in steady-state kinetics. Since normal sKIEs were obtained for all measurable acylation steps in the presteady state [ (D) k ac = 1.31 +/- 0.04, and the transient kinetic isotope effect at time zero (tKIE (0)) = 2.3 +/- 0.2], the kinetic step(s) contributing to the inverse sKIE of (D)( k cat/ K a) must occur more rapidly than the experimental time frame of the transient kinetics. Results are consistent with a chemical mechanism in which acylation occurs via a two-step process: the thiolate form of Cys-234, which is enriched in D 2O and gives rise to the inverse value of (D)( k cat/ K a), attacks the substrate to form a tetrahedral intermediate that proceeds to form an acyl-enzyme intermediate during a proton transfer step expressing a normal sKIE. The subsequent deacylation half-reaction is rate-limiting, with proton transfers exhibiting normal sKIEs. Through derivation of 12 equations describing all kinetic parameters and sKIEs for the proposed cathepsin C mechanism, integration of both steady-state and pre-steady-state kinetics with sKIEs allowed the provision of at least one self-consistent set of values for all 13 rate constants in this cysteine protease's chemical mechanism. Simulation of the resulting kinetic profile showed that at steady state approximately 80% of the enzyme exists in an active-site cysteine-acylated form in the mechanistic pathway. The chemical and kinetic details deduced from this work provide a potential roadmap to help steer drug discovery efforts for this and other disease-relevant cysteine proteases.  相似文献   

9.
The determination of kinetic parameters of chitinases using natural substrates is difficult due to low K(m) values, which require the use of low substrate concentrations that are hard to measure. Using the natural substrate (GlcNAc)(4), we have developed an assay for the determination of k(cat) and K(m)values of chitinases. Product concentrations as low as 0.5 microM were detected using normal-phase high-performance liquid chromatography (HPLC) with an amide 80 column (0.20 x 25 cm) using spectrophotometric detection at 210 nm. By means of this assay, k(cat) and K(m)values for chitinases A (ChiA) and B (ChiB) of Serratia marcescens were found to be 33+/-1s(-1) and 9+/-1 microM and 28+/-2s(-1) and 4+/-2 microM, respectively. For ChiB, these values were compared to those found with commonly used substrates where the leaving group is a (nonnatural) chromophore, revealing considerable differences. For example, assays with 4-methylumbelliferyl-(GlcNAc)(2) yielded a k(cat) value of 18+/-2s(-1) and a K(m) value of 30+/-6 microM. For two ChiB mutants containing a Trp --> Ala mutation in the +1 or +2 subsites, the natural substrate and the 4-methylumbelliferyl-(GlcNAc)(2) assays yielded rather similar K(m) values (5-fold difference at most) but showed dramatic differences in k(cat) values (up to 90-fold). These results illustrate the risk of using artificial substrates for characterization of chitinases and, thus, show that the new HPLC-based assay is a valuable tool for future chitinase research.  相似文献   

10.
NAD(+)-dependent 15-hydroxyprostaglandin dehydrogenase (15-PGDH), a member of the short-chain dehydrogenase/reductase (SDR) family, catalyzes the first step in the catabolic pathways of prostaglandins and lipoxins, and is believed to be the key enzyme responsible for the biological inactivation of these biologically potent eicosanoids. The enzyme utilizes NAD(+) specifically as a coenzyme. Potential amino acid residues involved in binding NAD(+) and facilitating enzyme catalysis have been partially identified. In this report, we propose that three more residues in 15-PGDH, Ile-17, Asn-91, and Val-186, are also involved in the interaction with NAD(+). Site-directed mutagenesis was used to examine their roles in binding NAD(+). Several mutants (I17A, I17V, I17L, I17E, I17K, N91A, N91D, N91K, V186A, V186I, V186D, and V186K) were prepared, expressed as glutathione S-transferase (GST) fusion enzymes in Escherichia coli, and purified by GSH-agarose affinity chromatography. Mutants I17E, I17K, N91L, N91K, and V186D were found to be inactive. Mutants N91A, N91D, V186A, and V186K exhibited comparable activities to the wild type enzyme. However, mutants I17A, I17V, I17L, and V186I had higher activity than the wild type. Especially, the activities of I17L and V186I were increased nearly 4- and 5-fold, respectively. The k(cat)/K(m) ratios of all active mutants for PGE(2) were similar to that of the wild type enzyme. However, the k(cat)/K(m) ratios of mutants I17A and N91A for NAD(+) were decreased 5- and 10-fold, respectively, whereas the k(cat)/K(m) ratios of mutants I17V, N91D, V186I, and V186K for NAD(+) were comparable to that of the wild type enzyme. The k(cat)/K(m) ratios of mutants I17L and V186A for NAD(+) were increased over nearly 2-fold. These results suggest that Ile-17, Asn-91, and Val-186 are involved in the interaction with NAD(+) and contribute to the full catalytic activity of 15-PGDH.  相似文献   

11.
The P450 2E1-catalyzed oxidation of ethanol to acetaldehyde is characterized by a kinetic deuterium isotope effect that increases K(m) with no effect on k(cat), and rate-limiting product release has been proposed to account for the lack of an isotope effect on k(cat) (Bell, L. C., and Guengerich, F. P. (1997) J. Biol. Chem. 272, 29643-29651). Acetaldehyde is also a substrate for P450 2E1 oxidation to acetic acid, and k(cat)/K(m) for this reaction is at least 1 order of magnitude greater than that for ethanol oxidation to acetaldehyde. Acetic acid accounts for 90% of the products generated from ethanol in a 10-min reaction, and the contribution of this second oxidation has been overlooked in many previous studies. The noncompetitive intermolecular kinetic hydrogen isotope effects on acetaldehyde oxidation to acetic acid ((H)(k(cat)/K(m))/(D)(k(cat)/K(m)) = 4.5, and (D)k(cat) = 1.5) are comparable with the isotope effects typically observed for ethanol oxidation to acetaldehyde, and k(cat) is similar for both reactions, suggesting a possible common catalytic mechanism. Rapid quench kinetic experiments indicate that acetic acid is formed rapidly from added acetaldehyde (approximately 450 min(-1)) with burst kinetics. Pulse-chase experiments reveal that, at a subsaturating concentration of ethanol, approximately 90% of the acetaldehyde intermediate is directly converted to acetic acid without dissociation from the enzyme active site. Competition experiments suggest that P450 2E1 binds acetic acid and acetaldehyde with relatively high K(d) values, which preclude simple tight binding as an explanation for rate-limiting product release. The existence of a rate-determining step between product formation and release is postulated. Also proposed is a conformational change in P450 2E1 occurring during the course of oxidation and the discrimination of P450 2E1 between acetaldehyde and its hydrated form, the gem-diol. This multistep P450 reaction is characterized by kinetic control of individual reaction steps and by loose binding of all ligands.  相似文献   

12.
Using oligonucleotide primers designed to the known gene sequence of an (E)-beta-farnesene (EbetaF) synthase, two cDNA sequences (MxpSS1 and MxpSS2) were cloned from a black peppermint (Menthaxpiperita) plant. MxpSS1 encoded a protein with 96% overall amino acid sequence identity with the EbetaF synthase. Recombinant MxpSS1 produced in Escherichia coli, after removal of an N-terminal thioredoxin fusion, had a K(m) for FPP of 1.91+/-0.1 microM and k(cat) of 0.18 s(-1), and converted farnesyl diphosphate (FPP) into four products, the major two being cis-muurola-3,5-diene (45%) and cis-muurola-4(14),5-diene (43%). This is the first cis-muuroladiene synthase, to be characterised. MxpSS2 encoded a protein with only two amino acids differing from EbetaF synthase. Recombinant MxpSS2 protein showed no activity towards FPP. One of the two mutations, at position 531 (leucine in MxpSS2 and serine in EbetaF synthase) was shown, by structural modelling to occur in the J-K loop, an element of the structure of sesquiterpene synthases known to be important in the reaction mechanism. Reintroduction of the serine at position 531 into MxpSS2 by site-directed mutagenesis restored EbetaF synthase activity (K(m) for FPP 0.98+/-0.12 microM, k(cat) 0.1 s(-1)), demonstrating the crucial role of this residue in the enzyme activity. Analysis, by GC-MS, of the sesquiterpene profile of the plant used for the cloning, revealed that EbetaF was not present, confirming that this particular mint chemotype had lost EbetaF synthase activity due to the observed mutations.  相似文献   

13.
Catalytic mechanism of hamster arylamine N-acetyltransferase 2   总被引:1,自引:0,他引:1  
Wang H  Liu L  Hanna PE  Wagner CR 《Biochemistry》2005,44(33):11295-11306
Arylamine N-acetyltransferases (NATs) catalyze an acetyl group transfer from AcCoA to primary arylamines, hydrazines, and hydrazides and play a very important role in the metabolism and bioactivation of drugs, carcinogens, and other xenobiotics. The reaction follows a ping-pong bi-bi mechanism. Structure analysis of bacterial NATs revealed a Cys-His-Asp catalytic triad that is strictly conserved in all known NATs. Previously, we have demonstrated by kinetic and isotope effect studies that acetylation of the hamster NAT2 is dependent on a thiolate-imidazolium ion pair (Cys-S(-)-His-ImH(+)) and not a general acid-base catalysis. In addition, we established that, after formation of the acetylated enzyme intermediate, the active-site imidazole, His-107, is likely deprotonated at physiological pH. In this paper, we report steady-state kinetic studies of NAT2 with two acetyl donors, acetyl coenzyme A (AcCoA) and p-nitrophenyl acetate (PNPA), and four arylamine substrates. The pH dependence of k(cat)/K(AcCoA) exhibited two inflection points at 5.32 +/- 0.13 and 8.48 +/- 0.24, respectively. The pK(a) at 5.32 is virtually identical with the previously reported pK(a) of 5.2 for enzyme acetylation, reaffirming that the first half of the reaction is catalyzed by a thiolate-imidazolium ion pair in the active site. The inflection point at 8.48 indicates that a pH-sensitive group on NAT2 is involved in AcCoA binding. A Br?nsted plot constructed by the correlation of log k(4) and log k(H)2(O) with the pK(a) for each arylamine substrate and water displays a linear free-energy relationship in the pK(a) range from -1.7 (H(2)O) to 4.67 (PABA), with a slope of beta(nuc) = 0.80 +/- 0.1. However, a further increase of the pK(a) from 4.67 (PABA) to 5.32 (anisidine) resulted in a 2.5-fold decrease in the k(4) value. Analysis of the pH-k(cat)/K(PABA) profile revealed a pK(a) of 5.52 +/- 0.14 and a solvent kinetic isotope effect (SKIE) of 2.01 +/- 0.04 on k(cat)/K(PABA). Normal solvent isotope effects of 4.8 +/- 0.1, 3.1 +/- 0.1, and 3.2 +/- 0.1 on the k(cat)/K(b) for anisidine, pABglu, and PNA, respectively, were also determined. These observations are consistent with a deacetylation mechanism dominated by nucleophilic attack of the thiol ester for arylamines with pK(a) values or=5.5. The general base is likely His-107 because the His-107 to Gln and Asn mutants were found to be devoid of catalytic activity. In contrast, an increase in pH-dependent hydrolysis of the acetylated enzyme was not observed over a pH range of 5.2-7.5. On the basis of these observations, a catalytic mechanism for the acetylation of arylamines by NAT2 is proposed.  相似文献   

14.
The activation of human platelets by alpha-thrombin is mediated at least in part by cleavage of protease-activated G-protein-coupled receptors, PAR-1 and PAR-4. Platelet glycoprotein Ibalpha also has a high affinity binding site for alpha-thrombin, and this interaction contributes to platelet activation through a still unknown mechanism. In the present study the hypothesis that GpIbalpha may contribute to platelet activation by modulating the hydrolysis of PAR-1 on the platelet membrane was investigated. Gel-filtered platelets from normal individuals were stimulated by alpha-thrombin, and the kinetics of PAR-1 hydrolysis by enzyme was followed with flow cytometry using an anti-PAR-1 monoclonal antibody (SPAN 12) that recognizes only intact PAR-1 molecules. This strategy allowed measurement of the apparent k(cat)/K(m) value for thrombin hydrolysis of PAR-1 on intact platelets, which was equal to 1.5 +/- 0.1 x 10(7) m(-1) sec(-1). The hydrolysis rate of PAR-1 by thrombin was measured under conditions in which thrombin binding to GpIb was inhibited by different strategies, with the following results. 1) Elimination of GpIbalpha on platelet membranes by mocarhagin treatment reduced the k(cat)/K(m) value by about 6-fold. 2) A monoclonal anti-GpIb antibody reduced the apparent k(cat)/K(m) value by about 5-fold. 3) An oligonucleotide DNA aptamer, HD22, which binds to the thrombin heparin-binding site (HBS) and inhibits thrombin interaction with GpIbalpha, reduced the apparent k(cat)/K(m) value by about 5-fold. 4) Displacement of alpha-thrombin from the binding site on GpIb using PPACK-thrombin reduced the apparent k(cat)/K(m) value by about 5-fold, and 5) mutation at the HBS of thrombin (R98A) caused a 5-fold reduction of the apparent k(cat)/K(m) value of PAR-1 hydrolysis. Altogether these results show that thrombin interaction with GpIb enhances the specificity of thrombin cleavage of PAR-1 on intact platelets, suggesting that GpIb may function as a "cofactor" for PAR-1 activation by thrombin.  相似文献   

15.
Human reticulocyte 15-lipoxygenase (15-hLO-1) and epithelial 15-lipoxygenase (15-hLO-2) have been implicated in a number of human diseases, with differences in their substrate specificity potentially playing a central role. In this paper, we present a novel method for accurately measuring the substrate specificity of the two 15-hLO isozymes and demonstrate that both cholate and specific LO products affect substrate specificity. The linoleic acid (LA) product, 13-hydroperoxyoctadienoic acid (13-HPODE), changes the ( k cat/ K m) (AA)/( k cat/ K m) (LA) ratio more than 5-fold for 15-hLO-1 and 3-fold for 15-hLO-2, while the arachidonic acid (AA) product, 12-( S)-hydroperoxyeicosatetraenoic acid (12-HPETE), affects only the ratio of 15-hLO-1 (more than 5-fold). In addition, the reduced products, 13-( S)-hydroxyoctadecadienoic acid (13-HODE) and 12-( S)-hydroxyeicosatetraenoic acid (12-HETE), also affect substrate specificity, indicating that iron oxidation is not responsible for the change in the ( k cat/ K m) (AA)/( k cat/ K m) (LA) ratio. These results, coupled with the dependence of the 15-hLO-1 k cat/ K m kinetic isotope effect ( (D) k cat/ K m) on the presence of 12-HPETE and 12-HETE, indicate that the allosteric site, previously identified in 15-hLO-1 [Mogul, R., Johansen, E., and Holman, T. R. (1999) Biochemistry 39, 4801-4807], is responsible for the change in substrate specificity. The ability of LO products to regulate substrate specificity may be relevant with respect to cancer progression and warrants further investigation into the role of this product-feedback loop in the cell.  相似文献   

16.
A new thermostable dipeptidase gene was cloned from the thermophile Brevibacillus borstelensis BCS-1 by genetic complementation of the D-Glu auxotroph Escherichia coli WM335 on a plate containing D-Ala-D-Glu. Nucleotide sequence analysis revealed that the gene included an open reading frame coding for a 307-amino-acid sequence with an M(r) of 35,000. The deduced amino acid sequence of the dipeptidase exhibited 52% similarity with the dipeptidase from Listeria monocytogenes. The enzyme was purified to homogeneity from recombinant E. coli WM335 harboring the dipeptidase gene from B. borstelensis BCS-1. Investigation of the enantioselectivity (E) to the P(1) and P(1)' site of Ala-Ala revealed that the ratio of the specificity constant (k(cat)/K(m)) for L-enantioselectivity to the P(1) site of Ala-Ala was 23.4 +/- 2.2 [E = (k(cat)/K(m))(L,D)/(k(cat)/K(m))(D,D)], while the D-enantioselectivity to the P(1)' site of Ala-Ala was 16.4 +/- 0.5 [E = (k(cat)/K(m))(L,D)/(k(cat)/K(m))(L,L)] at 55 degrees C. The enzyme was stable up to 55 degrees C, and the optimal pH and temperature were 8.5 and 65 degrees C, respectively. The enzyme was able to hydrolyze L-Asp-D-Ala, L-Asp-D-AlaOMe, Z-D-Ala-D-AlaOBzl, and Z-L-Asp-D-AlaOBzl, yet it could not hydrolyze D-Ala-L-Asp, D-Ala-L-Ala, D-AlaNH(2), and L-AlaNH(2.) The enzyme also exhibited beta-lactamase activity similar to that of a human renal dipeptidase. The dipeptidase successfully synthesized the precursor of the dipeptide sweetener Z-L-Asp-D-AlaOBzl.  相似文献   

17.
Lee SS  Yu S  Withers SG 《Biochemistry》2003,42(44):13081-13090
The unusual enzyme, Gracilariopsis alpha-1,4-glucan lyase of the sequence-related glycoside hydrolase family 31, cleaves the glycosidic bond of alpha-1,4-glucans via a beta-elimination reaction involving a covalent glycosyl-enzyme intermediate (Lee, S. S., Yu, S., and Withers, S. G. (2002) J. Am. Chem. Soc. 124, 4948-4949). The classical bell-shaped pH dependence of k(cat)/K(m) indicates two ionizable groups in the active site with apparent pK(a) values of 3.05 and 6.66. Br?nsted relationships of log k(cat) versus pK(a) and log(k(cat)/K(m)) versus pK(a) for a series of aryl glucosides both show a linear monotonic dependence on leaving group pK(a) with low beta(lg) values of 0.32 and 0.33, respectively. The combination of these low beta(lg) values with large secondary deuterium kinetic isotope effects (k(H)/k(D) = 1.16 - 1.19) on the first step indicate a glycosylation step with substantial glycosidic bond cleavage and proton donation to the leaving group oxygen at the transition state. Developed oxocarbenium ion character of the transition state is also suggested by the potent inhibition afforded by acarbose and 1-deoxynojirimycin (K(i) = 20 and 130 nM, respectively) and by the substantial rate reduction afforded by adjacent fluorine substitution. For only one substrate, 5-fluoro-alpha-D-glucopyranosyl fluoride, was the second elimination step shown to be rate-limiting. The large alpha-secondary deuterium kinetic isotope effect (k(H)/k(D) = 1.23) at C-1 and the small primary deuterium kinetic isotope effect (k(H)/k(D) = 1.92) at C-2 confirm an E2 mechanism with strong E1 character for this second step. This considerable structural and mechanistic similarity with retaining alpha-glucosidases is clear evidence for the evolution of an enzyme mechanism within the family.  相似文献   

18.
Chung HW  Tu SC 《Biochemistry》2012,51(24):4880-4887
Vibrio harveyi NADPH-FMN oxidoreductase (FRP) catalyzes flavin reduction by NADPH. In comparing amino acid sequence and crystal structure with Escherichia coli NfsA, residues N134, R225, R133, K167, and R15 were targeted for investigation of their possible roles in the binding and utilization of the NADPH substrate. By mutation of each of these five residues to an alanine, steady-state rate analyses showed that the variants K167A and R15A had apparently greatly increased K(m,NADPH) and reduced k(cat)/K(m,NADPH), whereas little or much more modest changes were found for the other variants. The deuterium isotope effects (D)(V/K) for (4R)-[4-(2)H]-NADPH were markedly increased to 6.3 and 7.4 for K167A and R15A, respectively, indicating that the rate constants for NADPH and NADP(+) dissociation were greatly enhanced relative to the hydride transfer steps. Also, anaerobic stopped-flow analyses revealed that the equilibrium dissociation constant for NADPH binding (K(d)) to be 2.5-3.9 and 1.1 mM for K167A and R15A, respectively, much higher than the 0.4 μM K(d) for the native FRP, whereas the k(cat) of these two variants were similar to that of the wild-type enzyme. Moreover, the K167 to alanine mutation led to even a slight increase in k(cat)/K(m) for NADH. These results, taken together, provide a strong support to the conclusion that K167 and R15 each was critical in the binding of NADPH by FRP. Such a functional role may also exist for other FRP homologous proteins.  相似文献   

19.
The kinetics for the isomerization of fructose-6-phosphate to glucose-6-phosphate (F6P --> G6P) by baker's yeast phosphoglucose isomerase (PGI) with regard to k(cat) and K(m) were determined from analysis of differential stopped flow microcalorimeter measurements using the integrated form of the Michaelis-Menten rate equation. Values for K(m) (F6P --> G6P) that were determined at pH 8.0 and ionic strength 0.1M at 293.4, 298.4, 303.4, and 311.5K exhibited a linear dependence on the substrate concentration at each temperature because of the substrate-product equilibrium. The minimum values for K(m) ranged from 2.62+/-0.55 mM at 293.4K to 7.8+/-4.8mM at 311.5K and were the same as the minimum values for the reverse reaction (G6P --> F6P) at 293.4 K and 298.4 K. Minimum values for k(cat) increased with temperature, from 2.78+/-0.34s(-1) at 293.4K to 11.4+/-1.0s(-1) at 311.5K, and for the reverse reaction, G6P --> F6P, from 0.852+/-0.086 s(-1) at 293.4K to 1.46+/-0.06s(-1) at 298.4K. The enzyme efficiency at 311.5K is close to the collision rate for a diffusion-controlled process in solution. The [F6P]/[G6P] equilibrium constants were determined from comparison of the values of k(cat) in both directions and were 0.307+/-0.053 at 293.4K and 0.395+/-0.033 at 298.4K. The heats of reaction in the F6P --> G6P direction increased from -8.96+/-0.26 kJmol(-1) at 311.5K to -8.27+/-0.40 kJmol(-1) at 293.4K, a value in fair agreement with 7.01+/-0.32 kJmol(-1) in the opposite G6P --> F6P direction.  相似文献   

20.
Murray BW  Padrique ES  Pinko C  McTigue MA 《Biochemistry》2001,40(34):10243-10253
Activation of receptor tyrosine kinases by autophosphorylation is one of the most common and critical transformations in signal transduction, yet its role in catalysis remains controversial. Autophosphorylation of the angiogenic receptor tyrosine kinase Tie2 was studied in terms of the autophosphorylation sites, sequence of phosphorylation at these sites, kinetic effects, and mechanistic consequences. Isoelectric focusing electrophoresis and mass spectrometric analysis of a Tie2 autophosphorylation time course showed that Tyr992 on the putative activation loop was phosphorylated first followed by Tyr1108 in the C-terminal tail (previously unidentified autophosphorylation site). Autophosphorylation of Tie2 to produce pTie2 resulted in a 100-fold increase in k(cat) and a 460-fold increase in k(cat)/K(m). Viscosity studies showed that the unphosphorylated Tie2 was partially limited by product diffusion ((k(cat))(eta) = 0.67 +/- 0.06), while product release was more rate-limiting ((k(cat))(eta) = 0.94 +/- 0.08) for autophosphorylated Tie2 (pTie2). Furthermore, autophosphorylation did not significantly affect the phosphoacceptor dissociation constants. There was a significant (k(cat))(H)/(k(cat))(D) solvent isotope effect (SIE) for unphosphorylated Tie2 (2.42 +/- 0.12) and modest SIE (1.28 +/- 0.04) for pTie2, which is consistent with the chemistry step being more rate-limiting for Tie2 as compared to pTie2. The pH-rate profiles of Tie2 and pTie2 revealed a >0.5 unit shift in the pK(a) values of catalytically relevant ionizable residues upon autophosphorylation. The shift in rate-limiting step will result in a different distribution of enzyme pools (e.g., E, E*S, E*P, etc.) which may modulate the susceptibility to inhibition. Tie2 and pTie2 were profiled with a panel of known ATP-competitive kinase inhibitors. Tie2 activation perturbs catalytic residue ionizations, shifts the rate-limiting step to almost exclusive diffusion-control, and transforms the kinase into a more perfect catalyst.  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