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.
M W Pinkse  M Merkx  B A Averill 《Biochemistry》1999,38(31):9926-9936
Purple acid phosphatases (PAPs) employ a dinuclear Fe(3+)Fe(2+) or Fe(3+)Zn(2+) center to catalyze the hydrolysis of phosphate monoesters. The interaction of fluoride with bovine spleen purple acid phosphatase (BSPAP) has been studied using a combination of steady-state kinetics and spectroscopic methods. For FeZn-BSPAP, the nature of the inhibition changes from noncompetitive at pH 6.5 (K(i(comp)) approximately K(i(uncomp)) approximately 2 mM) to uncompetitive at pH 5.0 (K(i(uncomp)) = 0.2 mM). The inhibition constant for AlZn-BSPAP at pH 5.0 (K(i) = 3 microM) is approximately 50-70-fold lower than that observed for both FeZn-BSAP and GaZn-BSPAP, suggesting that fluoride binds to the trivalent metal. Fluoride binding to the enzyme-substrate complex was found to be remarkably slow; hence, the kinetics of fluoride binding were studied in some detail for FeZn-, AlZn-, and FeFe-BSPAP at pH 5.0 and for FeZn-BSPAP at pH 6.5. Since the enzyme kinetics studies indicated the formation of a ternary enzyme-substrate-fluoride complex, the binding of fluoride to FeZn-BSPAP was studied using optical and EPR spectroscopies, both in the presence and absence of phosphate. The characteristic optical and EPR spectra of FeZn-BSPAP. F and FeZn-BSPAP.PO(4).F are similar at pH 5.0 and pH 6.5, indicating the formation of similar fluoride complexes at both pHs. A structural model for the ternary enzyme-(substrate/phosphate)-fluoride complexes is proposed that can explain the results from both the spectroscopic and the enzyme kinetics experiments. In this model, fluoride binds to the trivalent metal replacing the water/hydroxide ligand that is essential for the hydrolysis reaction to take place, while phosphate or the phosphate ester coordinates to the divalent metal ion.  相似文献   

2.
3.
4.
Ligand interactions with galactose oxidase: mechanistic insights.   总被引:1,自引:1,他引:0  
Interactions between galactose oxidase and small molecules have been explored using a combination of optical absorption, circular dichroism, and electron paramagnetic resonance (EPR) spectroscopies to detect complex formation and characterize the products. Anions bind directly to the cupric center in both active and inactive galactose oxidase, converting to complexes with optical and EPR spectra that are distinctly different from those of the starting aquo enzyme. Azide binding is coupled to stoichiometric proton uptake by the enzyme, reflecting the generation of a strong base (pKa > 9) in the active site anion adduct. At low temperature, the aquo enzyme converts to a form that exhibits the characteristic optical and EPR spectra of an anion complex, apparently reflecting deprotonation of the coordinated water. Anion binding results in a loss of the optical transition arising from coordinated tyrosine, implying displacement of the axial tyrosine ligand on forming the adduct. Nitric oxide binds to galactose oxidase, forming a specific complex exhibiting an unusual EPR spectrum with all g values below 2. The absence of Cu splitting in this spectrum and the observation that the cupric EPR signal from the active site metal ion is not significantly decreased in the complex suggest a nonmetal interaction site for NO in galactose oxidase. These results have been interpreted in terms of a mechanistic scheme where substrate binding displaces a tyrosinate ligand from the active site cupric ion, generating a base that may serve to deprotonate the coordinated hydroxyl group of the substrate, activating it for oxidation. The protein-NO interactions may probe a nonmetal O2 binding site in this enzyme.  相似文献   

5.
Initial velocity studies of rat liver cytosolic P-enolpyruvate carboxykinase in the direction of P-enolpyruvate formation gave intersecting double reciprocal plots indicating that the reaction conforms to a sequential reaction pathway. A complete product inhibition study with MnGDP-, P-enolpyruvate, and HCO3- as product inhibitors indicated that all patterns were noncompetitive. Isotope exchange at equilibrium with exchange between the substrate/product pairs GTP/GDP oxalacetate/HCO3-, and oxalacetate/P-enolpyruvate while varying the concentration of substrate/product pairs in fixed constant ratio gave no complete inhibitory patterns as the concentration of the constant ratio pairs approached saturation. The exchange rates between the substrate/product pairs differed by a factor of 40 when compared under the same assay conditions. These results were interpreted in terms of a random reaction mechanism in which true dead-end complexes do not form and in which the rate-limiting step is not the interconversion of the ternary quarternary central complexes. In addition to the formation of P-enolpyruvate from oxalacetate and MnGTP2-, the enzyme catalyzes the decarboxylation of oxalacetate to pyruvate in the absence of MnGTP2-. This reaction occurs only slowly in the absence of GDP and most rapidly in the presence of MnGDP-. When only MnGTP2- and oxalacetate are present, no pyruvate is formed, and oxalacetate is converted stoichiometrically to P-enolpyruvate. The enzyme also catalyzes the exchange of [14C]GDP into GTP in the absence of P-enolpyruvate. This exchange is stimulated by the presence of HCO3-. When enzyme is incubated with MnGTP2- in the presence or absence of HCO3-, there is no hydrolysis to form GDP and P1. The two partial reactions, namely the exchange of [14C]GDP with the E.HCO3.MnGTP or E.MnGTP complex and the formation of pyruvate from the E.oxalacetate.MnGDP complex provide pathways by which the expected dead-end complexes can be converted to enzyme forms which can return to the catalytic or exchange sequence.  相似文献   

6.
The mechanism of the L-threo-3-methylaspartate ammonia-lyase (EC 4.3.1.2) reaction has been probed using deuterium and solvent isotope effects with three different substrates, (2S,3S)-3-methylaspartic acid, (2S)-aspartic acid and (2S,3R)-3-methylaspartic acid. Each substrate appears to form a covalent adduct with the enzyme through the amination of a dehydroalanine (DehydAla-173) residue. The true substrates are N-protonated and at low pH, the alkylammonium groups are deprotonated internally in a closed solvent-excluded pocket after K+ ion, an essential cofactor, has become bound to the enzyme. At high pH, the amino groups of the substrates are able to react with the dehydroalanine residue prior to K+ ion binding. This property of the system gives rise to complex kinetics at pH 9.0 or greater and causes the formation of dead-end complexes which lack Mg2+ ion, another essential cofactor. The enzyme-substrate adduct is subsequently deaminated in two elimination processes. Hydrazines act as alternative substrates in the reverse reaction direction in the presence of fumaric acid derivatives, but cause irreversible inhibition in their absence. Borohydride and cyanide are not inhibitors. N-Ethylmaleimide also irreversibly inactivates the enzyme and labels residue Cys-361. The inactivation process is enhanced in the presence of cofactor Mg2+ ions and Cys-361 appears to serve as a base for the removal of the C-3 proton from the natural substrate, (2S,3S)-3-methylaspartic acid. The dehydroalanine residue appears to be protected in the resting form of the enzyme by generation of an internal thioether cross-link. The binding of the substrate and K+ ion appear to cause a conformational change which requires hydroxide ion. This is linked to reversal of the thioether protection step and generation of the base for substrate deprotonation at C-3. The deamination reaction displays high reverse reaction commitments and independent evidence from primary deuterium isotope effect data indicates that a thiolate acts as the base for deprotonation at C-3.  相似文献   

7.
Carboxypeptidase A: mechanism of zinc inhibition   总被引:1,自引:0,他引:1  
K S Larsen  D S Auld 《Biochemistry》1989,28(25):9620-9625
Zinc ions competitively inhibit carboxypeptidase A from bovine pancreas. The state(s) of hydroxylation of zinc and their possible site(s) of interaction with the enzyme have been investigated by determining the strength of zinc inhibition over pH range 4.6-10.5. The inhibition kinetics were recorded under stopped-flow conditions using the alpha-Val isozyme and the peptide substrate Dns-Gly-Ala-Phe in 0.5 M NaCl at 25 degrees C. The pH dependence of pKI follows a pattern which indicates that the enzyme is selectively inhibited by zinc monohydroxide, ZnOH+ (KI = 7.1 X 10(-7) M). The formation of the inhibitory ZnOH+ complex from fully hydrated Zn2+ is characterized by an ionization constant of 9.05, and the consecutive conversion of ZnOH+ to Zn(OH)2, Zn(OH)3-, and Zn(OH)4(2-) complexes takes place with ionization constants of 9.75, 10.1, and 10.5, respectively. Ionization of a ligand, LH, in the enzyme's inhibitory site (pKLH 5.8) is obligatory for binding of the ZnOH+ complex. The enzymatic activity (kcat/Km) is influenced by three ionizable groups: pKEH2 5.78, pKEH 8.60, and pKE 10.2. Since the values of pKLH and pKEH2 are virtually identical, it is possible that the inhibitory ZnOH+ complex interacts with the group responsible for pKEH2. Previous studies have suggested that pKEH2 reflects the ionization of Glu-270 and its interaction with a water molecule coordinated to the catalytic zinc ion. It is proposed that the inhibitory zinc ion binds to the carboxylate of Glu-270 and that the inhibition process is specific for zinc monohydroxide because it allows the formation of a stabilizing hydroxide bridge between the inhibitory and catalytic zinc ions.  相似文献   

8.
X Wu  B A Palfey  V V Mossine  V M Monnier 《Biochemistry》2001,40(43):12886-12895
Amadoriase is a flavoenzyme that catalyzes the oxidative deglycation of Amadori products (fructosyl amino acids or aliphatic amines) to yield free amine, glucosone, and hydrogen peroxide. The mechanism of action of amadoriase I from Aspergillus sp. has been investigated by stopped-flow kinetic studies using fructosyl propylamine and O(2) as substrates in 10 mM Tris HCl, pH 7.9, 4 degrees C. Using both substrate analogues and fast kinetic techniques, the active configuration of the substrate was found to be the beta-pyranose form. Stopped-flow studies showed that the reductive half-reaction is triphasic and generates intermediates that absorb at long wavelengths and is consistent either with (i) the reaction of the substrate with the flavin followed by iminium deprotonation or hydrolysis and then product release or with (ii) the formation of flavin reduction intermediates (carbanion equivalents or adducts), followed by product release. The rate of product release after flavin reduction is lower than the aerobic turnover rate, 14.4 s(-1), suggesting that it is not involved in the catalytic cycle and that reoxidation of the reduced enzyme occurs in the E(red)-product complex. In the oxidative half-reaction, the reduced flavin is oxidized by O(2) in a single phase. The observed rate constant has a linear dependence on oxygen concentration, giving a bimolecular rate constant of 4.9 x 10(4) M(-1) s(-1) in the absence of product, and 3.6 x 10(4) M(-1) s(-1) when the product is bound. The redox potentials of amadoriase have been measured at pH 7.0, 25 degrees, giving values of +48 and -52 mV for the oxidized enzyme/anionic semiquinone and anionic semiquinone/reduced enzyme couples, respectively.  相似文献   

9.
A divalent cation electrode was used to measure the stability constants (association constants) for the magnesium and manganese complexes of the substrates for the NADP+-specific isocitrate dehydrogenase (EC 1.1.1.42) from pea stems. At an ionic strength of 26.5 mM and at pH 7.4 the stability constants for the Mg2+-isocitrate and Mg2+-NADP+ complexes were 0.85 +/- 0.2 and 0.43 +/- 0.04 mM-1 respectively and for the Mn2+-isocitrate and Mn2+-NADP+ complexes they were 1.25 +/- 0.07 and 0.75 +/- 0.09 mM-1 respectively. At the same ionic strength but at pH 6.0 the Mg2+-NADPH and Mn2+-NADPH complexes had stability constants of 0.95 +/- 0.23 and 1.79 +/- 0.34 mM-1 respectively. Oxalosuccinate and alpha-ketoglutarate do not form measureable complexes under these conditions. Saturation kinetics of the enzyme with respect to isocitrate and metal ions are consistent with the metal-isocitrate complex being the substrate for the enzyme. NADP+ binds to the enzyme in the free form. Saturation kinetics of NADPH and Mn2+ indicate that the metal-NADPH complex is the substrate in the reverse reaction. In contrast the pig heart enzyme appears to bind free NADPH and Mn2+. A scheme for the reaction mechanism is presented and the difference between the reversibility of the NAD+ and NADP+ enzyme is discussed in relation to the stability of the NADH and NADPH metal complexes.  相似文献   

10.
Ribonucleotide reductase (RNR) is a key enzyme for the synthesis of the four DNA building blocks. Class Ia RNRs contain two subunits, denoted R1 (alpha) and R2 (beta). These enzymes are regulated via two nucleotide-binding allosteric sites on the R1 subunit, termed the specificity and overall activity sites. The specificity site binds ATP, dATP, dTTP, or dGTP and determines the substrate to be reduced, whereas the overall activity site binds dATP (inhibitor) or ATP. By using gas-phase electrophoretic mobility macromolecule analysis and enzyme assays, we found that the Escherichia coli class Ia RNR formed an inhibited alpha(4)beta(4) complex in the presence of dATP and an active alpha(2)beta(2) complex in the presence of ATP (main substrate: CDP), dTTP (substrate: GDP) or dGTP (substrate: ADP). The R1-R2 interaction was 30-50 times stronger in the alpha(4)beta(4) complex than in the alpha(2)beta(2) complex, which was in equilibrium with free alpha(2) and beta(2) subunits. Studies of a known E. coli R1 mutant (H59A) showed that deficient dATP inhibition correlated with reduced ability to form alpha(4)beta(4) complexes. ATP could also induce the formation of a generally inhibited alpha(4)beta(4) complex in the E. coli RNR but only when used in combination with high concentrations of the specificity site effectors, dTTP/dGTP. Both allosteric sites are therefore important for alpha(4)beta(4) formation and overall activity regulation. The E. coli RNR differs from the mammalian enzyme, which is stimulated by ATP also in combination with dGTP/dTTP and forms active and inactive alpha(6)beta(2) complexes.  相似文献   

11.
The visible absorption bands in the region 525-575 nm of the catalytic cobalt ion in cobalt(II) horse liver alcohol dehydrogenase show characteristic pH-dependent changes both in the free enzyme and its complexes with nicotinamide adenine dinucleotide (NAD+) and NAD+ plus ethanol or 2,2,2-trifluoroethanol. In the free enzyme, the change of the coordination environment has an apparent pK of about 9.4. In the binary complex with NAD+ the spectral changes are complex, indicating changes in the coordination sphere in a lower pH range with an estimated pK value of about 7.9. The ternary complexes enzyme X NAD+ X ethanol and enzyme X NAD+ X 2,2,2-trifluoroethanol exhibit very similar, characteristic spectral features; their apparent pK values are 6.3 and less than 4, respectively. We ascribe these pK values to the ionization of the alcohol bound in the ternary complexes. The results demonstrate that the catalytic cobalt ion is sensing changes of the ionization state of the protein when going from low pH forms to high pH forms both in the absence and presence of coenzyme and substrate/inhibitor.  相似文献   

12.
Tryptase, a serine protease with trypsin-like substrate cleavage properties, is one of the key effector molecules during allergic inflammation. It is stored in large quantities in the mast cell secretory granules in complex with heparin proteoglycan, and these complexes are released during mast cell degranulation. In the present paper, we have studied the mechanism for tryptase activation. Recombinant mouse tryptase, mouse mast cell protease 6 (mMCP-6), was produced in a mammalian expression system. The mMCP-6 fusion protein contained an N-terminal 6 x His tag followed by an enterokinase (EK) site replacing the native activation peptide (6xHis-EK-mMCP-6). In the absence of heparin, barely detectable enzyme activity was obtained after enterokinase cleavage of 6xHis-EK-mMCP-6 over a pH range of 5.5-7.5. However, when heparin was present, 6xHis-EK-mMCP-6 yielded active enzyme when enterokinase cleavage was performed at pH 5.5-6.0 but not at neutral pH. Affinity chromatography analysis showed that mMCP-6 bound strongly to heparin-Sepharose at pH 6.0 but not at neutral pH. After enterokinase cleavage of the sample at pH 6.0, mMCP-6 occurred in inactive monomeric form as shown by FPLC analysis on a Superdex 200 column. When heparin was added at pH 6.0, enzymatically active higher molecular weight complexes were formed, e.g., a dominant approximately 200 kDa complex that may correspond to tryptase tetramers. No formation of active tetramers was observed at neutral pH. When injected intraperitoneally, mMCP-6 together with heparin caused neutrophil influx, but no signs of inflammation were seen in the absence of heparin. The present paper thus indicates a crucial role for heparin in the formation of active mast cell tryptase.  相似文献   

13.
T S Angeles  G A Roberts  S A Carr  T D Meek 《Biochemistry》1992,31(47):11778-11784
The rates of desorption of the substrate water from the binary enzyme-H2O and ternary enzyme-H2O-(peptide)substrate complexes for the two hydrolases, porcine pepsin and thermolysin, have been investigated using a novel technique, solvent isotope partitioning. The experimental design of this method was based on the protocol of Rose et al. [Rose, I. A., O'Connell, E. L., Litwin, S., & BarTana, J. (1974) J. Biol. Chem. 249, 5163-5168] wherein the binary enzyme-H2(18)O complex established in the "pulse" solution was diluted into a "chase" solution containing variable concentrations of peptide substrates in a large pool of H2(16)O. The extent of trapping of H2(18)O within the respective E-H2(18)O and E-H2(18)O-(peptide)substrate complexes was determined from mass spectrometric analysis of the hydrolytic products. Our data have shown that the substrate water molecule of pepsin is not exclusively retained in the catalytic cycle and it desorbs from the apo- and substrate-bound complexes at rates that are at least 10 and 4 times faster, respectively, than that of product formation. Similarly, the low trapping of H2(18)O in the carboxylic product of the thermolysin reaction is a consequence of the ready desorption of H2(18)O from the ternary E-H2(18)O-(peptide)substrate complex and the binary E-H2(18)O complex. We attribute these results to the loss of the reactant water molecule due to desolvation of the enzyme's active site upon substrate binding.  相似文献   

14.
The specificity for substrate binding to creatine kinase for metal-nucleotide complexes of the type Cr-(H2O)4-n(NH3)nADP (where n = 0, 3, or 4) and Co-(H2O)4-m(NH3)mADP (for m = 3 or 4) has been investigated over the pH range 5.5-7.8 with the delta-alpha, beta-bidentate diastereoisomers. These inert nucleotide complexes acted as competitive inhibitors vs. MgADP over this range. In addition, the pH dependence of the V, V/K, and Km values for MgADP has been determined. Metal-nucleotide binding to the enzyme is strongest below an approximate pK of 6.45 but again becomes pH independent above pH 7. This pK is not associated with the metal-nucleotide complex. Instead, we conclude that the pK of the acid-base catalyst (thought to be histidine) is about 6.45 in the absence of nucleotide but is raised to 7.2 in its presence. This perturbation of the pK may result from a protein conformational change that allows a hydrogen bond to form between the phosphorylated nitrogen of phosphocreatine and the acid-base catalyst. The pK of the water in Cr(H2O)(NH3)3ADP has been determined to be 6.6, and by comparison of the binding affinity of this complex with that of Cr(NH3)4ADP or Cr(H2O)4ADP, it can be deduced that the hydroxo species binds more strongly than the aquo complex. In general, chromium nucleotides are bound more strongly than cobalt complexes, and binding affinity increases as water replaces ammonia in the first coordination sphere of the metal. Both trends are a result of stronger hydrogen-bond interactions between the metal complex and protein.  相似文献   

15.
The mechanism of sarcoplasmic reticulum (SR) ATPase Mg2+-dependent phosphorylation from Pi was investigated in the presence of 15% v/v dimethyl sulfoxide at pH 6, 20 degrees C, and in the absence of potassium. Measurements of intrinsic fluorescence changes and of 32P-labeled phosphoprotein (*E-P) were in agreement, both at equilibrium and in transient situations. We found that the amount of phosphoenzyme present and its rate of formation depended solely on the concentration of the (Mg X Pi) complex. Up to 6 nmol of phosphate/mg of protein was covalently bound to the enzyme, implying almost complete phosphorylation. Oxygen exchange experiments were also performed in order to allow calculation of the absolute rate constant of *E-P hydrolysis to the noncovalent complex (0.8-1.0 s-1), which differs from the observed rate of enzyme dephosphorylation (0.3-0.5 s-1); in addition, they allowed calculation of the bimolecular rate constant of substrate binding (2-2.4 M-1 s-1). The results demonstrate that in the presence of dimethyl sulfoxide, phosphorylation occurs by the following simple mechanism: relatively slow binding of the neutral substrate (Mg X Pi), with poor affinity, followed by a thermodynamically favorable formation of the covalent bond between phosphate and the possibly hydrophobic active site. The interaction between magnesium and calcium-deprived SR vesicles was studied in the presence of 0-20% v/v dimethyl sulfoxide (or 0-30% v/v glycerol) at pH 7 and 20 degrees C. The presence of either solvent led to the disappearance of the two typical pH-dependent effects we previously characterized for magnesium: loss of the Mg2+-induced spectral shift of tryptophan fluorescence emission and loss of the biphasic pattern displayed by the intrinsic fluorescence rise after addition of calcium to Ca2+-deprived Mg2+-preincubated vesicles. In the absence of solvent, the interaction of magnesium with the calcium-deprived ATPase was also characterized from the point of view of phosphoenzyme formation from ATP or Pi at pH 7 in the absence of potassium: we found that calcium-independent phosphorylation was slower when phosphate was added to SR vesicles preincubated with magnesium that when magnesium was added to vesicles preincubated with phosphate, suggesting that preincubation with magnesium had depleted the phosphate-reactive conformation of the ATPase. A simple reaction scheme for phosphoenzyme formation is described: it implies that the (Mg X Pi) complex is a substrate for this reaction, whereas the Mg2+ itself acts as a pH-dependent, dimethyl sulfoxide sensitive inhibitor of full enzyme phosphorylation.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

16.
1. The formation of the non-enzymic adduct of NAD(+) and sulphite was investigated. In agreement with others we conclude that the dianion of sulphite adds to NAD(+). 2. The formation of ternary complexes of either lactate dehydrogenase or malate dehydrogenase with NAD(+) and sulphite was investigated. The u.v. spectrum of the NAD-sulphite adduct was the same whether free or enzyme-bound at either pH6 or pH8. This suggests that the free and enzyme-bound adducts have a similar electronic structure. 3. The effect of pH on the concentration of NAD-sulphite bound to both enzymes was measured in a new titration apparatus. Unlike the non-enzymic adduct (where the stability change with pH simply reflects HSO(3) (-)=SO(3) (2-)+H(+)), the enzyme-bound adduct showed a bell-shaped pH-stability curve, which indicated that an enzyme side chain of pK=6.2 must be protonated for the complex to form. Since the adduct does not bind to the enzyme when histidine-195 of lactate dehydrogenase is ethoxycarbonylated we conclude that the protein group involved is histidine-195. 4. The pH-dependence of the formation of a ternary complex of lactate dehydrogenase, NAD(+) and oxalate suggested that an enzyme group is protonated when this complex forms. 5. The rate at which NAD(+) binds to lactate dehydrogenase and malate dehydrogenase was measured by trapping the enzyme-bound NAD(+) by rapid reaction with sulphite. The rate of NAD(+) dissociation from the enzymes was calculated from the bimolecular association kinetic constant and from the equilibrium binding constant and was in both cases much faster than the forward V(max.). No kinetic evidence was found that suggested that there were interactions between protein subunits on binding NAD(+).  相似文献   

17.
Methylglyoxal (MG) was studied as an inhibitor and effective glycating factor of human muscle-specific enolase. The inhibition was carried out by the use of a preincubation procedure in the absence of substrate. Experiments were performed in anionic and cationic buffers and showed that inhibition of enolase by methylglyoxal and formation of enolase-derived glycation products arose more effectively in slight alkaline conditions and in the presence of inorganic phosphate. Incubation of 15 micromolar solutions of the enzyme with 2 mM, 3.1 mM and 4.34 mM MG in 100 mM phosphate buffer pH 7.4 for 3 h caused the loss a 32%, 55% and 82% of initial specific activity, respectively. The effect of MG on catalytic properties of enolase was investigated. The enzyme changed the K(M) value for glycolytic substrate 2-phospho-D-glycerate (2-PGA) from 0.2 mM for native enzyme to 0.66 mM in the presence of MG. The affinity of enolase for gluconeogenic substrate phosphoenolpyruvate altered after preincubation with MG in the same manner, but less intensively. MG has no effect on V(max) and optimal pH values. Incubation of enolase with MG for 0-48 h generated high molecular weight protein derivatives. Advanced glycation end products (AGEs) were resistant to proteolytic degradation by trypsin. Magnesium ions enhanced the enzyme inactivation by MG and facilitated AGEs formation. However, the protection for this inhibition in the presence of 2-PGA as glycolytic substrate was observed and AGEs were less effectively formed under these conditions.  相似文献   

18.
Xu H  West AH  Cook PF 《Biochemistry》2006,45(39):12156-12166
Kinetic data have been measured for the histidine-tagged saccharopine dehydrogenase from Saccharomyces cerevisiae, suggesting the ordered addition of nicotinamide adenine dinucleotide (NAD) followed by saccharopine in the physiologic reaction direction. In the opposite direction, the reduced nicotinamide adenine dinucleotide (NADH) adds to the enzyme first, while there is no preference for the order of binding of alpha-ketoglutarate (alpha-Kg) and lysine. In the direction of saccharopine formation, data also suggest that, at high concentrations, lysine inhibits the reaction by binding to free enzyme. In addition, uncompetitive substrate inhibition by alpha-Kg and double inhibition by NAD and alpha-Kg suggest the existence of an abortive E:NAD:alpha-Kg complex. Product inhibition by saccharopine is uncompetitive versus NADH, suggesting a practical irreversibility of the reaction at pH 7.0 in agreement with the overall K(eq). Saccharopine is noncompetitive versus lysine or alpha-Kg, suggesting the existence of both E:NADH:saccharopine and E:NAD:saccharopine complexes. NAD is competitive versus NADH, and noncompetitive versus lysine and alpha-Kg, indicating the combination of the dinucleotides with free enzyme. Dead-end inhibition studies are also consistent with the random addition of alpha-Kg and lysine. Leucine and oxalylglycine serve as lysine and alpha-Kg dead-end analogues, respectively, and are uncompetitive against NADH and noncompetitive against alpha-Kg and lysine, respectively. Oxaloacetate (OAA), pyruvate, and glutarate behave as dead-end analogues of lysine, which suggests that the lysine-binding site has a higher affinity for keto acid analogues than does the alpha-Kg site or that dicarboxylic acids have more than one binding mode on the enzyme. In addition, OAA and glutarate also bind to free enzyme as does lysine at high concentrations. Glutarate gives S-parabolic noncompetitive inhibition versus NADH, indicating the formation of a E:(glutarate)2 complex as a result of occupying both the lysine- and alpha-Kg-binding sites. Pyruvate, a slow alternative keto acid substrate, exhibits competitive inhibition versus both lysine and alpha-Kg, suggesting the combination to the E:NADH:alpha-Kg and E:NADH:lysine enzyme forms. The equilibrium constant for the reaction has been measured at pH 7.0 as 3.9 x 10(-7) M by monitoring the change in NADH upon the addition of the enzyme. The Haldane relationship is in very good agreement with the directly measured value.  相似文献   

19.
Steady state kinetic studies have been performed to investigate the formation of thrombin from prothrombin by human coagulation Factor Xa in the presence of Ca2+ and phospholipid. The concentration of ligand which gives 50% of the maximum velocity (K0.5) is 2.3 mM for Ca2+, 7.4 microM for phospholipid, and 0.006 microM for prothrombin. Hill plots of the Ca2+ enhancement of the reaction give a Hill coefficient of 3.1, indicating positive cooperativity. The initial velocity patterns are consistent with an ordered addition of reactants with phospholipid as the second reactant to bind to the enzyme. Although our results do not differentiate between Ca2+ or the prothrombin substrate as the first reactant to bind to Factor Xa, it is established that Ca2+ can bind to Factor Xa in the absence of the other reactants. Thus, the most probable order of addition of reactants is Ca2+, phospholipid, and the prothrombin substrate. Plots of (v)-1 versus (prothrombin)-1 or (v)-1 versus [(Ca2+)3]-1 at several constant concentrations of phospholipid indicate that the major effect of phospholipid is to increase the turnover number of Factor Xa.  相似文献   

20.
A novel colanic acid-degrading enzyme was isolated from a mixed culture filtrate obtained by enrichment culturing of a compost sample using colanic acid as carbon source. The enzyme was partially purified resulting in a 17-fold increase in specific activity. Further purification by Native PAGE revealed that the enzyme is part of a high-molecular weight multi protein complex of at least six individual proteins. The enzyme showed a temperature optimum at 50 degrees C while after 5h at 50 degrees C and pH7 still 70% of the total activity was left. The pH optimum was found to be pH7. Analysis of the degradation products showed that the enzyme is a novel 1,4-beta-fucoside hydrolase that liberates repeating units of colanic acid with varying degrees of acetylation. Km and Vmax of the enzyme were determined against the native substrate as well as its de-O-acetylated and depyruvated forms. Compared to the native substrate the affinity of the enzyme for the modified substrates was much lower. However, for the de-O-acetylated sample a dramatic increase in catalytic efficiency was observed. The native form of the substrate showed substrate inhibition at high concentrations, probably due to the formation of nonproductive substrate complexes. Removal of the acetyl groups probably prevents this effect resulting in a higher catalytic efficiency.  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