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 468 毫秒
1.
1. The curved plots of 1/v against 1/[S] obtained when caeruloplasmin oxidizes NN-dimethyl-p-phenylenediamine were investigated. The first free-radical oxidation product of caeruloplasmin oxidation of NN-dimethyl-p-phenylenediamine is required for curvature, as straight-line plots were obtained when activities were measured either before appreciable free-radical product had appeared or in the presence of ascorbate, which reduced it back to NN-dimethyl-p-phenylenediamine. 2. In the presence of ascorbate linear reciprocal-plots were obtained with all of the 37 substrates tested. V(max.) values varied over only an eightfold range and those for the 20 p-amino compounds over only a twofold range. K(m) values, however, varied over a 10(4)-fold range. The small range of V(max.) values indicates that the rate-limiting step in caeruloplasmin action is relatively independent of the nature of the substrate. K(m) values suggest that substrates bind primarily by ring electrons, although certain side-chain groups increased the K(m) in a manner unrelated to likely changes of ring-electron densities. A mechanism involving repulsion between negative charges on the substrate and the enzyme was supported by the variation of the K(m) of 5-hydroxyindol-3-ylacetic acid with pH.  相似文献   

2.
Kinetic measurements were made with cortisone reductase (20-dihydrocortisone-NAD(+) oxidoreductase, EC 1.1.1.53) and a series of substrates which differed in shape, size and electronic character in the region adjacent to C-11, C-14 and C-18. Structural changes at C-11 in these substrates resulted in up to 660-fold changes in the apparent K(m) value, up to 200-fold changes in the apparent V(max.) value and up to 800-fold changes in the ratio of these kinetic constants. It is suggested that interactions important for substrate function normally occur between the enzyme and the C ring in the region of C-11, that these interactions arise from so-called hydrophobic forces between the generally hydrophobic C ring portion of the substrate and a hydrophobic region of the enzyme, but that when the substrate contains a polar substituent in this portion of the molecule, then polar interactions with polar moieties of the enzyme can also be important. It is further suggested that the part of the enzyme that interacts with the region of C-11 in the substrate is flexible, and that substrate binding involves at least some degree of induced fit.  相似文献   

3.
The cDNA for the alpha-isoenzyme from rhesus monkey (Macaca mulatta) liver was cloned and expressed in yeast. The alpha-isoenzymes of human and monkey liver alcohol dehydrogenase differ from the other human and horse liver enzymes in having Met57, Ala93, and Val116 instead of Leu57, Phe93, and Leu116 in the substrate binding pocket and Gly47 instead of Arg47 near the pyrophosphate moiety of the coenzyme. The effects of these differences on the kinetic mechanism, substrate specificity, and coenzyme binding were studied with the purified, recombinant monkey alpha-isoenzyme (MmADH alpha) and mutated enzymes with Gly47 substituted with His or Arg. The mechanism appears to be random for the binding of NAD+ and ethanol and ordered for NADH and acetaldehyde, with formation of a dead-end enzyme-NADH-ethanol complex. MmADH alpha reacts 130-fold slower (V/K) with ethanol and 3-25-fold slower with 2-methyl alcohols but 20-fold faster with cyclohexanol, as compared with horse (Equus caballus) liver EE isoenzyme (EqADH). MmADH alpha is stereoselective for the R isomer of 2-butanol, whereas EqADH favors the S isomer. Both enzymes have comparable reactivity with larger primary alcohols. MmADH alpha is more reactive with secondary alcohols and has highest activity with cyclohexanol. However, it does not react with steroids such as 5 beta-androstane-17 beta-ol-3-one. Molecular modeling suggests that the differences between MmADH alpha and EqADH are a result of the substitution of Ala for Phe93 and Thr for Ser48. MmADH alpha binds NAD+ most rapidly when a group with a pK of 7.4 is unprotonated, implicating His51 in this reaction. The G47R substitution decreased the dissociation constants for NAD+ and NADH and turnover numbers only about 2-fold, whereas the G47H substitution increased dissociation constants 7-14-fold and turnover numbers 4-fold. A basic residue at position 47 is not crucial for activity, as multiple interactions determine coenzyme affinity.  相似文献   

4.
The parkinsonian inducing agent, 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP), is a cyclic tertiary allylamine exhibiting good monoamine oxidase B (MAO-B) substrate properties. MAO-B catalyzes the ring alpha-carbon 2-electron bioactivation of MPTP to yield the 1-methyl-4-phenyl-2,3-dihydropyridinium species (MPDP(+)). The corresponding 5-membered ring MPTP analogue, 1-methyl-3-phenyl-3-pyrroline, also undergoes MAO-B-catalyzed oxidation to give the 2-electron oxidation product, 1-methyl-3-phenylpyrrole. Here we report the kinetic deuterium isotope effects on V(max) and V(max)/K(m) for the steady-state oxidation of 1-methyl-3-phenyl-3-pyrroline and 1-methyl-3-(4-fluorophenyl)-3-pyrroline by baboon liver MAO-B, using the corresponding pyrroline-2,2,4,5,5-d(5) analogues as the deuterated substrates. The apparent isotope effects for the two substrates were 4.29 and 3.98 on V(max), while the isotope effects on V(max)/K(m) were found to be 5.71 and 3.37, respectively. The values reported for the oxidation of MPTP by bovine liver MAO-B with MPTP-6,6-d(2), as deuterated substrate, are (D)(V(max))=3.55; (D)(V(max)/K(m))=8.01. We conclude that the mechanism of the MAO-B-catalyzed oxidation of pyrrolinyl substrates is similar to that of the tetrahydropyridinyl substrates and that a carbon-hydrogen bond cleavage step is, at least partially, rate determining.  相似文献   

5.
Genetic defects in pyruvate dehydrogenase complex (PDC) cause lactic acidosis, neurological deficits, and often early death. Most mutations of PDC are localized in the alpha subunit of the pyruvate dehydrogenase (E1) component. We have kinetically characterized a patient's missense mutation alphaH44R in E1alpha by creating and purifying three recombinant human E1s (alphaH44R, alphaH44Q, and alphaH44A). Substitutions at histidine-15 resulted in decreased V(max) values (6% alphaH44R; 30% alphaH44Q; 90% alphaH44A) while increasing K(m) values for thiamine pyrophosphate (TPP) compared to wild-type (alphaH44R, 3-fold; alphaH44Q, 7-fold; alphaH44A, 10-fold). This suggests that the volume of the residue at site 15 is important for TPP binding and substitution by a residue with a longer side chain disrupts the active site more than the TPP binding site. The rates of phosphorylation and dephosphorylation of alphaH44R E1 by E1-kinase and phospho-E1 phosphatase, respectively, were similar to that of the wild-type E1 protein. These results provide a biochemical basis for altered E1 function in the alphaH44R E1 patient.  相似文献   

6.
The roles of Phe-120 and Glu-222 in the oxidation of chiral substrates bunitrolol (BTL) and bufuralol (BF) by CYP2D6 are discussed. Wild-type CYP2D6 (CYP2D6-WT) oxidized BTL to 4-hydroxybunitrolol (4-OH-BTL) with substrate enantioselectivity of (R)-(+)-BTL > (S)-(-)-BTL. The same enzyme converted BF into 1'-hydroxybufuralol with substrate enantioselectivity of (R)-BF > (S)-BF and metabolite diastereoselectivity of (1'R)-OH < (1'S)-OH. The substitution of Phe-120 by alanine markedly increased the apparent K(m) and V(max) values for enantiomeric BTL 4-hydroxylation by CYP2D6. In contrast, the same substitution caused an increase only in V(max) values of (S)-BF 1'-hydroxylation without changing apparent K(m) values, while kinetic parameters (K(m) and V(max) values) for (R)-BF 1'-hydroxylation remained unchanged. Furthermore, the substitution of Glu-222 as well as Glu-216 by alanine remarkably decreased both the apparent K(m) and V(max) values without changing substrate enantioselectivity or metabolite diastereoselectivity. A computer-assisted simulation study using energy minimization and molecular dynamics techniques indicated that the hydrophobic interaction of an aromatic moiety of the substrate with Phe-120 and the ionic interaction of a basic nitrogen atom of the substrate with Glu-222 in combination with Glu-216 play important roles in the binding of BF and BTL by CYP2D6 and the orientation of these substrates in the active-site cavity. This modeling yielded a convincing explanation for the reversal of substrate enantioselectivity in BTL 4-hydroxylation between CYP2D6-WT and CYP2D6-V374M having methionine in place of Val-374, which supports the validity of this modeling.  相似文献   

7.
Sharma R  Deng H  Leung A  Mahuran D 《Biochemistry》2001,40(18):5440-5446
In humans, beta-hexosaminidase A (alphabeta) is required to hydrolyze GM2 ganglioside. A deficiency of either the alpha- or beta-subunit leads to a severe neurological disease, Tay-Sachs or Sandhoff disease, respectively. In mammals beta-hexosaminidase B (betabeta) and S (alphaalpha) are other major and minor isozymes. The primary structures of the alpha- and beta-subunits are 60% identical, but only the alpha-containing isozymes can efficiently hydrolyze beta-linked GlcNAc-6-SO(4) from natural or artificial substrates. Hexosaminidase has been grouped with glycosidases in family 20. A molecular model of the active site of the human hexosaminidase has been generated from the crystal structure of a family 20 bacterial chitobiase. We now use the chitobiase structure to identify residues close to the carbon-6 oxygen of NAG-A, the nonreducing beta-GlcNAc residue of its bound substrate. The chitobiase side chains in the best interactive positions align with alpha-Asn(423)Arg(424) and beta-Asp(453)Leu(454). The change in charge from positive in alpha to negative in beta is consistent with the lower K(m) of hexosaminidase S, and the much higher K(m) and lower pH optimum of hexosaminidase B, toward sulfated versus unsulfated substrates. In vitro mutagenesis, CHO cell expression, and kinetic analyses of an alphaArg(424)Lys hexosaminidase S detected little change in V(max) but a 2-fold increase in K(m) for the sulfated substrate. Its K(m) for the nonsulfated substrate was unaffected. When alphaAsn(423) was converted to Asp, again only the K(m) for the sulfated substrate was changed, increasing by 6-fold. Neutralization of the charge on alphaArg(424) by substituting Gln produced a hexosaminidase S with a K(m) decrease of 3-fold and a V(max) increased by 6-fold for the unsulfated substrate, parameters nearly identical to those of hexosaminidase B at pH 4.2. As well, for the sulfated substrate at pH 4.2 its K(m) was increased 9-fold and its V(max) decreased 1.5-fold, values very similar to those of hexosaminidase B obtained at pH 3.0, where its betaAsp(453) becomes protonated.  相似文献   

8.
Recombinant human cytosolic sialidase (HsNEU2), expressed in Escherichia coli, was purified to homogeneity, and its substrate specificity was studied. HsNEU2 hydrolyzed 4-methylumbelliferyl alpha-NeuAc, alpha 2-->3 sialyllactose, glycoproteins (fetuin, alpha-acid glycoprotein, transferrin, and bovine submaxillary gland mucin), micellar gangliosides GD1a, GD1b, GT1b, and alpha 2-->3 paragloboside, and vesicular GM3. alpha 2-->6 sialyllactose, colominic acid, GM1 oligosaccharide, whereas micellar GM2 and GM1 were resistant. The optimal pH was 5.6, kinetics Michaelis-Menten type, V(max) varying from 250 IU/mg protein (GD1a) to 0.7 IU/mg protein (alpha(1)-acid glycoprotein), and K(m) in the millimolar range. HsNEU2 was activated by detergents (Triton X-100) only with gangliosidic substrates; the change of GM3 from vesicular to mixed micellar aggregation led to a 8.5-fold V(max) increase. HsNEU2 acted on gangliosides (GD1a, GM1, and GM2) at nanomolar concentrations. With these dispersions (studied in detailed on GM1), where monomers are bound to the tube wall or dilutedly associated (1:2000, mol/mol) to Triton X-100 micelles, the V(max) values were 25 and 72 microIU/mg protein, and K(m) was 10 and 15 x 10(-9) m, respectively. Remarkably, GM1 and GM2 were recognized only as monomers. HsNEU2 worked at pH 7.0 with an efficiency (compared with that at pH 5.6) ranging from 4% (on GD1a) to 64% (on alpha(1)-acid glycoprotein), from 7% (on GD1a) to 45% (on GM3) in the presence of Triton X-100, and from 30 to 40% on GM1 monomeric dispersion. These results show that HsNEU2 differentially recognizes the type of sialosyl linkage, the aglycone part of the substrate, and the supramolecular organization (monomer/micelle/vesicle) of gangliosides. The last ability might be relevant in sialidase interactions with gangliosides under physiological conditions.  相似文献   

9.
The CCR4 family proteins are 3'-5'-deadenylases that function in the first step of the degradation of poly(A) mRNA. Here we report the purification to homogeneity of the yeast CCR4 protein and the analysis of its substrate specificities. CCR4 deadenylated a 7N+23A substrate (seven nucleotides followed by 23 A residues) in a distributive manner. Only small differences in CCR4 activity for different A length substrates were observed until only 1 A residue remained. Correspondingly, the K(m) for a 25N+2A substrate was found to be at least 20-fold lower than that for a 26N+1A substrate, although their V(max) values differed by only 2-fold. In addition, the total length of the RNA was found to contribute to CCR4 activity: up to 17 nucleotides (not necessarily poly(A)) could be recognized by CCR4. Poly(U), poly(C), and poly(G) were also found to be 12-30-fold better inhibitors of CCR4 compared with poly(A), supporting the observation that CCR4 contains a non-poly(A)-specific binding site. Surprisingly, even longer substrates (>/=45 nucleotides) stimulated CCR4 to become a processive enzyme, suggesting that CCR4 undergoes an additional transition in the presence of such substrates. CCR4 also displayed no difference in its activity with capped or uncapped RNA substrates. These results indicate that CCR4 recognition of its RNA substrates involves several features of the RNA that could be sites in vivo for controlling the rate of specific mRNA deadenylation.  相似文献   

10.
1. The effects of varying pH, ionic strength and temperature on the parameters K(m) and V(max.) for a purified alkaline phosphatase from calf intestinal mucosa with a new fluorogenic substrate, 4-methylumbelliferyl phosphate monoester disodium salt, and an ammediol-hydrochloric acid buffer system were determined. 2. It was found that, under varying conditions, a relationship exists between K(m) and V(max.) such that V(max.)=beta/(1+alpha/K(m)), where alpha and beta are constants, temperature- and ionic strength-dependent, but pH-independent. It is shown that this relationship accounts satisfactorily for the well-known effect of varying substrate concentration on optimum pH and velocity. 3. The various results are interpreted in terms of a pH-dependent conformational equilibrium between two forms of the enzyme, E(1) and E(2). Only E(1) combines with substrate, and only E(2) reacts to give inorganic phosphate. 4. To account for the pH-variation of K(m) and V(max.) in terms of this theory, it is postulated that the conformational change is associated with a change in pK of two basic groups in the enzyme.  相似文献   

11.
A genetic variant of PNPLA3 (patatin-like phospholipase domain-containing 3; PNPLA3-I148M), a serine protease of unknown function, is associated with accumulation of triacylglycerol (TAG) in the liver. To determine the biological substrates of PNPLA3 and the effect of the I148M substitution on enzymatic activity and substrate specificity, we purified and characterized recombinant human PNPLA3 and PNPLA3-I148M. Maximal hydrolytic activity of PNPLA3 was observed against the three major glycerolipids, TAG, diacylglycerol, and monoacylglycerol, with a strong preference for oleic acid as the acyl moiety. Substitution of methionine for isoleucine at position 148 markedly decreased the V(max) of the enzyme for glycerolipids but had only a modest effect on the K(m). Purified PNPLA3 also catalyzed the hydrolysis of oleoyl-CoA, but the V(max) was 100-fold lower for oleoyl-CoA than for triolein. The thioesterase activity required the catalytic serine but was only modestly decreased by the I148M substitution. The enzyme had little or no hydrolytic activity against the other lipid substrates tested, including phospholipids, cholesteryl ester, and retinyl esters. Neither the wild-type nor mutant enzyme catalyzed transfer of oleic acid from oleoyl-CoA to glycerophosphate, lysophosphatidic acid, or diacylglycerol, suggesting that the enzyme does not promote de novo TAG synthesis. Taken together, our results are consistent with the notion that PNPLA3 plays a role in the hydrolysis of glycerolipids and that the I148M substitution causes a loss of function, although we cannot exclude the possibility that the enzyme has additional substrates or activities.  相似文献   

12.
The retro steroids 17beta-hydroxy-5beta,9beta,10alpha-androstan-3-one and 5beta,9beta,10alpha-androstane-3,17-dione were good substrates for cortisone reductase in the presence of NADH, and the products corresponded to the respective 3beta-hydroxy compounds, in which the 3beta-hydroxyl group is axial and the absolute configuration is 3S. The analogous natural steroids 17beta-hydroxy-5beta,9alpha,10beta-androstan-3-one and 5beta,9alpha,10beta-androstane-3,17-dione were very poor substrates, and gave the corresponding 3alpha(equatorial,3R)-hydroxy compounds, and, in the latter case, also an appreciable amount of 3beta(axial, 3S)-hydroxy-5beta,9alpha,10beta-androstan-17-one. The natural steroids 17beta-hydroxy-5alpha,9alpha,10beta-androstan-3-one and 5alpha,9alpha,10beta-androstane-3,17-dione were better substrates than the retro steroid 17beta-hydroxy-5alpha,9beta,10alpha-androstan-3-one, but were not such good substrates as the retro steroids 17beta-hydroxy-5beta,9beta,10alpha-androstan-3-one and 5beta,9beta,10alpha-androstane-3,17-dione. Unlike these retro steroid 5beta,9beta,10alpha-androstan-3-ones, the natural steroids 17beta-hydroxy-5alpha,9alpha,10beta-androstan-3-one and 5alpha,9alpha,10beta-androstane-3,17-dione gave the corresponding 3alpha(axial,3R)-hydroxy compounds. The retro steroid 17beta-hydroxy-5alpha,9beta,10alpha-androstan-3-one was not a good substrate, and the product of reaction corresponded to the 3alpha(axial,3R)-hydroxy compound. The nature of substrate recognition by this enzyme is discussed in the light of these structure-activity relationships.  相似文献   

13.
Two forms of alpha-galactosidase, I and II, exist in Vicia faba seeds and these have been purified 3660- and 337-fold respectively. They behaved as homogeneous preparations when examined by ultracentrifugation, disc electrophoresis and gel filtration. The apparent molecular weights of enzymes I and II, as determined by gel filtration, were 209000 and 38000 respectively. The carbohydrate contents of enzymes I and II were 25% and 2.8% respectively, and the enzymes differed in their aromatic amino acid compositions. Enzyme I was split into six inactive subunits in the presence of 6m-urea. alpha-Galactosidases I and II showed different pH optima and K(m) and V(max.) values with p-nitrophenyl alpha-d-galactoside and raffinose as substrates, and also differed in their thermal stabilities.  相似文献   

14.
Analysis of SCE frequencies in Chinese hamster V79 cells was used to investigate the influence of the stereoisomeric forms of epoxides in mammalian genotoxicity tests. The SCE-inducing potency of 12 pairs of (R)- and (S)-enantiomeric epoxides which differed in the degree of substitution of the oxirane ring was determined. Of these, 2 pairs of epoxides failed to induce SCE. Different SCE-inducing potencies between the (R)- and (S)-enantiomers were shown for 5 epoxides. This study demonstrates that stereoselectivity might play an important role in genotoxicity testing of chemicals with asymmetric C atoms.  相似文献   

15.
1. The deoxyfluoro-d-glucopyranose 6-phosphates are substrates for both yeast and rat liver glucose 6-phosphate dehydrogenase. 2. The V(max.) values (relative to d-glucose 6-phosphate) were determined for a series of d-glucose 6-phosphate derivatives substituted at C-2. The V(max.) values decreased with increasing electronegativity of the C-2 substituent. This is consistent with a mechanism involving hydride-ion transfer. 3. 2-Deoxy-d-arabino-hexose 6-phosphate (2-deoxy-d-glucose 6-phosphate) showed substrate inhibition with the yeast enzyme but not with the rat liver enzyme. 4. 2-Amino-2-deoxy-d-glucose 6-phosphate (d-glucosamine 6-phosphate) was a substrate for the yeast enzyme but a competitive inhibitor for the rat liver enzyme. 5. Lineweaver-Burk plots for the d-glucose 6-phosphate derivatives with yeast glucose 6-phosphate dehydrogenase were biphasic.  相似文献   

16.
1. Three fractions of beta-galactosidase activity from the rat small-intestinal mucosa were separated chromatographically. Two of these fractions had an acid pH optimum at 3-4, and the third one had a more neutral pH optimum at 5.7. 2. The two ;acid' beta-galactosidase fractions had considerably lower K(m) values for hetero beta-galactosides than for lactose. The V(max.) values were similar for all the substrates used (lactose, phenyl beta-galactoside, o-nitrophenyl beta-galactoside, p-nitrophenyl beta-galactoside and 6-bromo-2-naphthyl beta-galactoside). No difference could be detected between the two ;acid' fractions with respect to their enzymic properties (pH optimum, K(m) for the different substrates, K(i) for lactose as an inhibitor of the hydrolysis of hetero beta-galactosides, K(i) for phenyl beta-galactoside as an inhibitor of the hydrolysis of lactose, and relative V(max.) for the hydrolysis of different substrates). These two fractions probably represent different forms of the same enzyme. 3. The ;neutral' fraction had similar K(m) values for all the substrates hydrolysed, but with lactose as substrate the V(max.) was much higher than with the hetero beta-galactosides. This fraction did not split phenyl beta-galactoside or 6-bromo-2-naphthyl beta-galactoside at a measurable rate. 4. Lactose was a competitive inhibitor of the hetero beta-galactosidase activities of all the three fractions, and K(i) for lactose as an inhibitor in each case was the same as K(m) for the lactase activity. Phenyl beta-galactoside was a competitive inhibitor of the lactase activity of all the three fractions. These facts strongly indicate that in all the three fractions lactose is hydrolysed by the same active sites as the hetero beta-galactosides. 5. Human serum albumin stabilized the separated enzymes against inactivation by freezing and thawing.  相似文献   

17.
Fructokinase from beef liver showed a clear reversal in specificity when the two isomers of ATP beta S were used as substrates with Mg2+ and Cd2+, with the Sp isomer having the higher V/K value with Mg2+ and the Rp isomer the higher value with Cd2+. The delta isomer of MgATP is thus the active form of the substrate. The substitution of sulfur for oxygen in the noncoordinated position of the beta-phosphate caused a 102-fold decrease in V/K over the value seen with MgATP, while substitution in the coordinated position gave a 21-fold decrease over the V/K value seen with CdATP. The Km values were little affected by sulfur substitution, showing that the wrong screw sense isomers were nonproductively bound almost as well as the correct ones. When ADP alpha S was used as a substrate in the reverse reaction, the Sp isomer showed the highest V/K value with both Mg2+ and Cd2+, suggesting that the metal ion is not coordinated to the alpha-phosphate during transphosphorylation. The failure of CrATP to act as a substrate for fructokinase suggests that the enzyme inserts one of its side chains into the inner coordination sphere of the metal ion during the reaction.  相似文献   

18.
The relationship between the structure and activity of meta- and para-hydroxylated monophenols was studied during their tyrosinase-catalysed hydroxylation and the rate-limiting steps of the reaction mechanism were identified. The para-hydroxylated substrates permit us to study the effect of a substituent (R) in the carbon-1 position (C-1) of the benzene ring on the nucleophilic attack step, while the meta group permits a similar study of the effect on the electrophilic attack step. Substrates with a -OCH3 group on C-1, as p-hydroxyanisol (4HA) and m-hydroxyanisol (3HA), or with a -CH2OH group, as p-hydroxybenzylalcohol (4HBA) and m-hydroxybenzylalcohol (3HBA), were used because the effect of the substituent (R) size was assumed to be similar. However, the electron-donating effect of the -OCH3 group means that the carbon-4 position (C-4) is favoured for nucleophilic attack (para-hydroxylated substrates) or for electrophilic attack (meta-hydroxylated substrates). The electron-attracting effect of the -CH2OH group has the opposite effect, hindering nucleophilic (para) or electrophilic (meta) attack of C-4. The experimental data point to differences between the maximum steady-state rate (V(M)Max) of the different substrates, the value of this parameter depends on the nucleophilic and electrophilic attack. However, differences are greatest in the Michaelis constants (K(M)m), with the meta-hydroxylated substrates having very large values. The catalytic efficiency k(M)cat/K(M)m is much greater for thepara-hydroxylated substrates although it varies greatly between one substrate and the other. However, it varies much less in the meta-hydroxylated substrates since this parameter describes the power of the nucleophilic attack, which is weaker in the meta OH. The large increase in the K(M)m of the meta-hydroxylated substrates might suggest that the phenolic OH takes part in substrate binding. Since this is a weaker nucleophil than the para-hydroxylated substrates, the binding constant decreases, leading to an increase in K(M)m. The catalytic efficiency of tyrosinase on a monophenol (para or meta) is directly related to the nucleophilic power of the oxygen of the phenolic OH. The oxidation step is not limiting since if this were the case, the para and meta substrates would have the same V(M)max. The small difference between the absolute values of V(M)max suggests that the rate constants of the nucleophilic and electrophilic attacks are on the same order of magnitude.  相似文献   

19.
The E and S isoenzymes of horse liver alcohol dehydrogenase differ by 10 amino acid residues, but only the S isoenzyme is active on 3 beta-hydroxysteroids. This functional difference was correlated to the differences in structures of the isoenzymes by characterizing a series of chimeric enzymes, which could represent intermediates in the evolution of catalytic activity. Deletion of Asp-115 from the E isoenzyme created the E/D115 delta enzyme that is active on steroids. The deletion alters the substrate binding pocket by moving Leu-116, which sterically hinders binding of steroids in the E isoenzyme. A chimeric enzyme (ESE) that has four changes in or near the substrate binding pocket (T94I/R101S/F110L/D115 delta) was 15-30-fold more catalytically efficient (V/Km) on uncharged steroids than was the E/D115 delta enzyme. Molecular modeling suggests that the substitutions at residues 94 and 110 indirectly affect the activity on steroids. ESE enzyme was 6-fold more active than the S isoenzyme on neutral steroids, due to substitutions not in the substrate binding pocket. The K366E and the Q17E/A43T/A59T substitutions in the S isoenzyme gave 2-fold increases in V/Km on steroids, which together can account for the changes observed with the ESE enzyme. The enzymes that are active on steroids did not bind 2,2,2-trifluoroethanol as tightly and were catalytically less efficient than the E isoenzyme with small alcohols. However, these enzymes were two to three and four to five orders of magnitude more efficient with 1-hexanol and 5 beta-androstane-3 beta,17 beta-diol, respectively, than with ethanol. These results demonstrate that several residues not directly participating in substrate binding or chemical catalysis contribute to catalytic efficiency.  相似文献   

20.
Mevalonate kinase serine/threonine residues have been implicated in substrate binding and inherited metabolic disease. Alignment of >20 mevalonate kinase sequences indicates that Ser-145, Ser-146, Ser-201, and Thr-243 are the only invariant residues with alcohol side chains. These residues have been individually mutated to alanine. Structural integrity of the mutants has been demonstrated by binding studies using fluorescent and spin-labeled ATP analogs. Kinetic characterization of the mutants indicates only modest changes in K(m)((ATP)). K(m) for mevalonate increases by approximately 20-fold for S146A, approximately 40-fold for T243A, and 100-fold for S201A. V(max) changes for S145A, S201A, and T243A are < or =3-fold. Thus, the 65-fold activity decrease associated with the inherited human T243I mutation seems attributable to the nonconservative substitution rather than any critical catalytic function. V(max) for S146A is diminished by 4000-fold. In terms of V/K(MVA), this substitution produces a 10(5)-fold effect, suggesting an active site location and catalytic role for Ser-146. The large k(cat) effect suggests that Ser-146 productively orients ATP during catalysis. K(D(Mg-ATP)) increases by almost 40-fold for S146A, indicating a specific role for Ser-146 in liganding Mg(2+)-ATP. Instead of mapping within a proposed C-terminal ATP binding motif, Ser-146 is situated in a centrally located motif, which characterizes the galactokinase/homoserine kinase/ mevalonate kinase/phosphomevalonate kinase protein family. These observations represent the first functional demonstration that this region is part of the active site in these related phosphotransferases.  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