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 46 毫秒
1.
Bacterial phosphotriesterase (PTE) catalyzes the hydrolysis of a wide variety of organophosphate nerve agents and insecticides. Previous kinetic studies with a series of enantiomeric organophosphate triesters have shown that the wild type PTE generally prefers the S(P)-enantiomer over the corresponding R(P)-enantiomers by factors ranging from 1 to 90. The three-dimensional crystal structure of PTE with a bound substrate analogue has led to the identification of three hydrophobic binding pockets. To delineate the factors that govern the reactivity and stereoselectivity of PTE, the dimensions of these three subsites have been systematically altered by site-directed mutagenesis of Cys-59, Gly-60, Ser-61, Ile-106, Trp-131, Phe-132, His-254, His-257, Leu-271, Leu-303, Phe-306, Ser-308, Tyr-309, and Met-317. These studies have shown that substitution of Gly-60 with an alanine within the small subsite dramatically decreased k(cat) and k(cat)/K(a) for the R(P)-enantiomers, but had little influence on the kinetic constants for the S(P)-enantiomers of the chiral substrates. As a result, the chiral preference for the S(P)-enantiomers was greatly enhanced. For example, the value of k(cat)/K(a) with the mutant G60A for the S(P)-enantiomer of methyl phenyl p-nitrophenyl phosphate was 13000-fold greater than that for the corresponding R(P)-enantiomer. The mutation of I106, F132, or S308 to an alanine residue, which enlarges the small or leaving group subsites, caused a significant reduction in the enantiomeric preference for the S(P)-enantiomers, due to selective increases in the reaction rates for the R(P)-enantiomers. Enlargement of the large subsite by the construction of an H254A, H257A, L271A, or M317A mutant had a relatively small effect on k(cat)/K(a) for either the R(P)- or S(P)-enantiomers and thus had little effect on the overall stereoselectivity. These studies demonstrate that by modifying specific residues located within the active site of PTE, it is possible to dramatically alter the stereoselectivity and overall reactivity of the native enzyme toward chiral substrates.  相似文献   

2.
Organophosphorus acid anhydrolase (OPAA) catalyzes the hydrolysis of p-nitrophenyl analogs of the organophosphonate nerve agents, sarin and soman. The enzyme is stereoselective toward the chiral phosphorus center by displaying a preference for the R(P)-configuration of these analogs. OPAA also exhibits an additional preference for the stereochemical configuration at the chiral carbon center of the soman analog. The preferred configuration of the chiral carbon center is dependent upon the configuration at the phosphorus center. The enzyme displays a two- to four-fold preference for the R(P)-enantiomer of the sarin analog. The k(cat)/K(m) of the R(P)-enantiomer is 250 M(-1) s(-1), while that of the S(P)-enantiomer is 110 M(-1) s(-1). The order of preference for the stereoisomers of the soman analog is R(P)S(C) > R(P)R(C) > S(P)R(C) > S(P)S(C). The k(cat)/K(m) values are 36,300 M(-1)s(-1), 1250 M(-1) s(-1), 80 M(-1) s(-1) and 5 M(-1) s(-1), respectively. The R(P)S(C)-isomer of the soman analog is therefore preferred by a factor of 7000 over the S(P)S(C)-isomer.  相似文献   

3.
Hardy LW  Kirsch JF 《Biochemistry》1984,23(6):1275-1282
The Bacillus cereus beta-lactamase I catalyzes the hydrolysis of a wide variety of penicillins and cephalosporins with values of k(cat)/K(m) varying over several orders of magnitude. The values of this parameter for the most reactive of these compounds, benzylpenicillin, I, and furylacryloyl-penicillin, II (k(cat)/K(m) = 2.43 x 10(7) M(-1) s(-1) and 2.35 x 10(7) M(-1) s(-1), respectively, at pH 7.0 in potassium phosphate buffer containing 0.17 M KCl, I(c) = 0.63, 25 degrees C) are decreased markedly by increasing viscosity in sucrose- or glycerol-containing buffers. The relative sensitivities to viscosity of k(cat)/K(m) values for I and for cephaloridine, III, were found to be virtually unchanged at pH 3.8 from those observed at pH 7.0. The differential effects of viscosity on the reactive vs. the sluggish [e.g., cephalothin (IV), k(cat)/K(m) = 1 x 10(4) M(-1) s(-1)] substrates support the contention that the rates of reaction of the former with the enzyme are in part diffusion controlled. Quantitative analysis gives values for the association rate constants, k(1), of 7.6 x 10(7) M(-1) s(-1), 4 x 10(7) M(-1) s(-1), and 1.1 x 10(7) M(-1) s(-1) for I, II, and III, respectively. As both reactive and sluggish substrates associate with the active site of the enzyme with relatively similar rate constants, the variation in k(cat)/K(m) values is primarily due to the variation in the partition ratios k(-1)/k(2), for the ES complex, which are 2.3, 0.77, and 30 for I, II, and III, respectively. The preceding analysis is based on direct application of the Stokes-Einstein diffusion law to enzyme kinetics. The range of applicability of this law to the diffusion of substrate size molecules and the mechanics of diffusion of ionic species through viscous solutions of sucrose vs. polymers are explored.  相似文献   

4.
Extracellular Penicillium fellutanum exo-beta-D-galactofuranosidase, with a mass of 70 kDa, was purified to apparent homogeneity. The enzyme was used to investigate the influence of phosphodiesters of the peptidophosphogalactomannans pP(2)GM(ii) and pP(25)GM(ii) (containing 2 and 25 phosphodiester residues, respectively, per mol of polymer) on the kinetic parameters of galactofuranosyl hydrolysis of these two polymers, of 1-O-methyl-beta-D-galactofuranoside, and of two galactofuranooligosaccharides. The enzyme did not hydrolyze phosphorylated galactose residues of pP(2)GM(ii) or pP(25)GM(ii). The k(cat)/K(m) value for pP(25)GM(ii) is 1.7 x 10(3) M(-1) s(-1), that for 1-O-methyl-beta-D-galactofuranoside is 1.1 x 10(4) M(-1) s(-1), that for pP(2)GM(ii) is 1.7 x 10 (4) M(-1) s(-1), and those for 5-O-beta-D-galactofuranooligosaccharides with degrees of polymerization of 3.4 and 5.5 are 1.7 x 10(5) and 4.1 x 10(5) M(-1) s(-1), respectively. Variability in the k(cat)/K(m) values is due primarily to differences in K(m) values; the k(-1)/k(1) ratio likely provides the most influence on K(m). k(cat) increases as the degree of polymerization of galactofuranosyl residues increases. Most of the galactofuranosyl and phosphocholine residues were removed by day 8 in vivo from pP(x)GM(ii) added to day 3 cultures initiated in medium containing 2 mM phosphate but not from those initially containing 20 mM phosphate. The filtrates from day 9 cultures initiated in 2 mM inorganic phosphate in modified Raulin-Thom medium contained 0.2 mM inorganic phosphate and 2.2 U of galactofuranosidase ml(-1)h(-1). No galactofuranosidase activity but 15 mM inorganic phosphate was found in filtrates from day 9 cultures initiated in 20 mM phosphate. In vivo the rate of galactofuranosyl hydrolysis of pP(x)GM(ii) and of related polymers is proportional to the k(cat)/K(m) value of each polymer. The kinetic data show that the k(cat)/K(m) value increases as the number of phosphodiesters of pP(x)GM(ii) decreases, also resulting in an increase in the activity of exo-beta-D-galactofuranosidase.  相似文献   

5.
Cdc25A is a dual-specific protein phosphatase involved in the regulation of the kinase activity of Cdk-cyclin complexes in the eukaryotic cell cycle. To understand the mechanism of this important regulator, we have generated highly purified biochemical reagents to determine the kinetic constants for human Cdc25A with respect to a set of peptidic, artificial, and natural substrates. Cdc25A and its catalytic domain (dN25A) demonstrate very similar kinetics toward the artificial substrates p-nitrophenyl phosphate (k(cat)/K(m) = 15-25 M(-1) s(-1)) and 3-O-methylfluorescein phosphate (k(cat)/K(m) = 1.1-1.3 x 10(4) M(-1) s(-1)). Phospho-peptide substrates exhibit extremely low second-order rate constants and a flat specificity profile toward Cdc25A and dN25A (k(cat)/K(m) = 1 to 10 M(-1) s(-1)). In contrast to peptidic substrates, Cdc25A and dN25A are highly active phosphatases toward the natural substrate, T14- and Y15-bis-phosphorylated Cdk2/CycA complex (Cdk2-pTpY/CycA) with k(cat)/K(m) values of 1.0-1.1 x 10(6) M(-1) s(-1). In the context of the Cdk2-pTpY/CycA complex, phospho-threonine is preferred over phospho-tyrosine by more than 10-fold. The highly homologous catalytic domain of Cdc25c is essentially inactive toward Cdk2-pTpY/CycA. Taken together these data indicate that a significant degree of the specificity of Cdc25 toward its Cdk substrate resides within the catalytic domain itself and yet is in a region(s) that is outside the phosphate binding site of the enzyme.  相似文献   

6.
The Escherichia coli protein YjeQ represents a protein family whose members are broadly conserved in bacteria and have been shown to be indispensable to the growth of E. coli and Bacillus subtilis [Arigoni, F., et al. (1998) Nat. Biotechnol. 16, 851]. Proteins of the YjeQ family contain all sequence motifs typical of the vast class of P-loop-containing GTPases, but show a circular permutation, with a G4-G1-G3 pattern of motifs as opposed to the regular G1-G3-G4 pattern seen in most GTPases. All YjeQ family proteins display a unique domain architecture, which includes a predicted N-terminal OB-fold RNA-binding domain, the central permuted GTPase module, and a zinc knuckle-like C-terminal cysteine cluster. This domain architecture suggests a possible role for YjeQ as a regulator of translation. YjeQ was overexpressed, purified to homogeneity, and shown to contain 0.6 equiv of GDP. Steady state kinetic analyses indicated slow GTP hydrolysis, with a k(cat) of 9.4 h(-)(1) and a K(m) for GTP of 120 microM (k(cat)/K(m) = 21.7 M(-)(1) s(-)(1)). YjeQ also hydrolyzed other nucleoside triphosphates and deoxynucleotide triphosphates such as ATP, ITP, and CTP with specificity constants (k(cat)/K(m)) ranging from 0.2 to 1.0 M(-)(1) s(-)(1). Pre-steady state kinetic analysis of YjeQ revealed a burst of nucleotide hydrolysis for GTP described by a first-order rate constant of 100 s(-)(1) as compared to a burst rate of 0.2 s(-)(1) for ATP. In addition, a variant in the G1 motif of YjeQ (S221A) was substantially impaired for GTP hydrolysis (0.3 s(-)(1)) with a less significant impact on the steady state rate (1.8 h(-)(1)). In summary, E. coli YjeQ is an unusual, circularly permuted P-loop-containing GTPase, which catalyzes GTP hydrolysis at a rate 45 000 times greater than that of turnover.  相似文献   

7.
The ability to redesign enzymes to catalyze noncognate chemical transformations would have wide-ranging applications. We developed a computational method for repurposing the reactivity of metalloenzyme active site functional groups to catalyze new reactions. Using this method, we engineered a zinc-containing mouse adenosine deaminase to catalyze the hydrolysis of a model organophosphate with a catalytic efficiency (k(cat)/K(m)) of ~10(4) M(-1) s(-1) after directed evolution. In the high-resolution crystal structure of the enzyme, all but one of the designed residues adopt the designed conformation. The designed enzyme efficiently catalyzes the hydrolysis of the R(P) isomer of a coumarinyl analog of the nerve agent cyclosarin, and it shows marked substrate selectivity for coumarinyl leaving groups. Computational redesign of native enzyme active sites complements directed evolution methods and offers a general approach for exploring their untapped catalytic potential for new reactivities.  相似文献   

8.
Chlorophyllase catalyzes the initial step in the degradation of chlorophyll and plays a key role in leaf senescence and fruit ripening. Here, we report the cloning of chlorophyllase from Triticum aestivum (wheat) and provide a detailed mechanistic analysis of the enzyme. Purification of recombinant chlorophyllase from an Escherichia coli expression system indicates that the enzyme functions as a dimeric protein. Wheat chlorophyllase hydrolyzed the phytol moiety from chlorophyll (k(cat) = 566 min(-1); K(m) = 63 microM) and was active over a broad temperature range (10-75 degrees C). In addition, the enzyme displays carboxylesterase activity toward p-nitrophenyl (PNP)-butyrate, PNP-decanoate, and PNP-palmitate. The pH-dependence of the reaction showed the involvement of an active site residue with a pK(a) of approximately 6.5 for both k(cat) and k(cat)/K(m) with chlorophyll, PNP-butyrate, and PNP-decanoate. Using these substrates, solvent kinetic isotope effects ranging from 1.5 to 1.9 and from 1.4 to 1.9 on k(cat) and k(cat)/K(m), respectively, were observed. Proton inventory experiments suggest the transfer of a single proton in the rate-limiting step. Our analysis of wheat chlorophyllase indicates that the enzyme uses a charge-relay mechanism similar to other carboxylesterases for catalysis. Understanding the activity and mechanism of chlorophyllase provides insight on the biological and chemical control of senescence in plants and lays the groundwork for biotechnological improvement of this enzyme.  相似文献   

9.
In this study, we have shown that recombinant BH1999 from Bacillus halodurans catalyzes the hydrolysis of gentisyl coenzyme A (CoA) (2,5-dihydroxybenzoyl-coenzyme A) at physiological pH with a k(cat)/K(m) of 1.6 x 10(6) M(-1) s(-1) and the hydrolysis of 3-hydroxybenzoyl-CoA with a k(cat)/K(m) of 3.0 x 10(5) M(-1) s(-1). All other acyl-CoA thioesters tested had low or no substrate activity. The BH1999 gene is juxtaposed with a gene cluster that contains genes believed to function in gentisate oxidative degradation. It is hypothesized that BH1999 functions as a gentisyl-CoA thioesterase. Gentisyl-CoA thioesterase shares the backbone fold and the use of an active site aspartate residue to mediate catalysis with the 4-hydroxybenzoyl-CoA thioesterase of the hotdog fold enzyme superfamily. A comparative study of these two enzymes showed that they differ greatly in the rate contribution made by the catalytic aspartate, in the pH dependence of catalysis, and in substrate specificity.  相似文献   

10.
11.
An intramolecularly quenched fluorogenic peptide containing o-aminobenzoyl (Abz) and ethylenediamine 2,4-dinitrophenyl (Eddnp) groups at amino- and carboxyl-terminal amino acid residues, Abz-Lys-Pro-Ile-Glu-Phe-Phe-Arg-Leu-Eddnp, was hydrolyzed by purified human pepsin, gastricsin, and gastric juice uniquely at the Phe-Phe bond. Kinetic parameters determined for purified pepsin were K(m)=0.68+/-0.11 microM; k(cat)=6.3+/-0.16s(-1); k(cat)/K(m)=9.26s(-1) microM(-1); Gastricsin showed K(m)=2.69+/-0.18 microM; k(cat)=0.03+/-0.005s(-1); k(cat)/K(m)=0.011s(-1) microM(-1). Gastric juice (21 samples) from subjects without gastric disorders at endoscopy examination showed activities varying from 0.0008 to 9.72 micromolml(-1)min(-1). Pepstatin A inhibition of gastric juice enzymatic activity was complete at 3.4x10(-5)M (final concentration) inhibitor. In the proposed method the presence of a unique scissile bond in the synthetic substrate provides a direct ratio between enzymatic activity and amount of substrate hydrolyzed, and a unique step reaction facilitates the use of this assay for the determination of the activity of aspartic proteinases in biological fluids and during enzyme purification procedures.  相似文献   

12.
It has been suggested that the mechanism of alkaline phosphatase (AP) is associative, or triester-like, because phosphorothioate monoesters are hydrolyzed by AP approximately 10(2)-fold slower than phosphate monoesters. This "thio effect" is similar to that observed for the nonenzymatic hydrolysis of phosphate triesters, and is the inverse of that observed for the nonenzymatic hydrolysis of phosphate monoesters. The latter reactions proceed by loose, dissociative transition states, in contrast to reactions of triesters, which have tight, associative transition states. Wild-type alkaline phosphatase catalyzes the hydrolysis of p-nitrophenyl phosphate approximately 70 times faster than p-nitrophenyl phosphorothioate. In contrast, the R166A mutant alkaline phosphatase enzyme, in which the active site arginine at position 166 is replaced with an alanine, hydrolyzes p-nitrophenyl phosphate only about 3 times faster than p-nitrophenyl phosphorothioate. Despite this approximately 23-fold change in the magnitude of the thio effects, the magnitudes of Bronsted beta(lg) for the native AP (-0.77 +/- 0.09) and the R166A mutant (-0.78 +/- 0. 06) are the same. The identical values for the beta(lg) indicate that the transition states are similar for the reactions catalyzed by the wild-type and the R166A mutant enzymes. The fact that a significant change in the thio effect is not accompanied by a change in the beta(lg) indicates that the thio effect is not a reliable reporter for the transition state of the enzymatic phosphoryl transfer reaction. This result has important implications for the interpretation of thio effects in enzymatic reactions.  相似文献   

13.
The beta-xylosidase from Aspergillus awamori X-100 belonging to the family 3 glycoside hydrolase revealed a distinctive transglycosylating ability to produce xylooligosaccharides with degree of polymerization more than 7. In order to explain this fact, the enzyme has been subjected to the detailed biochemical study. The enzymatic hydrolysis of p-nitrophenyl beta-D-xylopyranoside was found to occur with overall retention of substrate anomeric configuration suggesting cleavage of xylosidic bonds through a double-displacement mechanism. Kinetic study with aryl beta-xylopyranosides substrates, in which leaving group pK(a)s were in the range of 3.96-10.32, revealed monotonic function of log(k(cat)) and no correlation of log(k(cat)/Km) versus pKa values indicating deglycosylation as a rate-limiting step for the enzymatic hydrolysis. The classical bell-shaped pH dependence of k(cat)/Km indicated two ionizable groups in the beta-xylosidase active site with apparent pKa values of 2.2 and 6.4. The kinetic parameters of hydrolysis, Km and k(cat), of p-nitrophenyl beta-D-1,4-xylooligosaccharides were very close to those for hydrolysis of p-nitrophenyl-beta-D-xylopyranoside. Increase of p-nitrophenyl-beta-D-xylopyranoside concentration up to 80 mM led to increasing of the reaction velocity resulting in k(cat)(app)=81 s(-1). Addition of alpha-methyl D-xylopyranoside to the reaction mixture at high concentration of p-nitrophenyl-beta-D-xylopyranoside (50 mM) caused an acceleration of the beta-xylosidase-catalyzed reactions and appearance of a new transglycosylation product, alpha-methyl D-xylopyranosyl-1,4-beta-D-xylopyranoside, that was identified by 1H NMR spectroscopy. The kinetic model suggested for the enzymatic reaction was consistent with the results obtained.  相似文献   

14.
Human angiotensin-converting enzyme-related carboxypeptidase (ACE2) is a zinc metalloprotease whose closest homolog is angiotensin I-converting enzyme. To begin to elucidate the physiological role of ACE2, ACE2 was purified, and its catalytic activity was characterized. ACE2 proteolytic activity has a pH optimum of 6.5 and is enhanced by monovalent anions, which is consistent with the activity of ACE. ACE2 activity is increased approximately 10-fold by Cl(-) and F(-) but is unaffected by Br(-). ACE2 was screened for hydrolytic activity against a panel of 126 biological peptides, using liquid chromatography-mass spectrometry detection. Eleven of the peptides were hydrolyzed by ACE2, and in each case, the proteolytic activity resulted in removal of the C-terminal residue only. ACE2 hydrolyzes three of the peptides with high catalytic efficiency: angiotensin II () (k(cat)/K(m) = 1.9 x 10(6) m(-1) s(-1)), apelin-13 (k(cat)/K(m) = 2.1 x 10(6) m(-1) s(-1)), and dynorphin A 1-13 (k(cat)/K(m) = 3.1 x 10(6) m(-1) s(-1)). The ACE2 catalytic efficiency is 400-fold higher with angiotensin II () as a substrate than with angiotensin I (). ACE2 also efficiently hydrolyzes des-Arg(9)-bradykinin (k(cat)/K(m) = 1.3 x 10(5) m(-1) s(-1)), but it does not hydrolyze bradykinin. An alignment of the ACE2 peptide substrates reveals a consensus sequence of: Pro-X((1-3 residues))-Pro-Hydrophobic, where hydrolysis occurs between proline and the hydrophobic amino acid.  相似文献   

15.
Neurolysin (EP24.16) and thimet oligopeptidase (EP24.15) are closely related metalloendopeptidases. Site-directed mutagenesis of Tyr(613) (EP24.16) or Tyr(612) (EP24.15) to either Phe or Ala promoted a strong reduction of k(cat)/K(M) for both enzymes. These data suggest the importance of both hydroxyl group and aromatic ring at this specific position during substrate hydrolysis by these peptidases. Furthermore, the EP24.15 A607G mutant showed a k(cat)/K(M) of 2x10(5) M(-1) s(-1) for the Abz-GFSIFRQ-EDDnp substrate, similar to that of EP24.16 (k(cat)/K(M)=3x10(5) M(-1) s(-1)) which contains Gly at the corresponding position; the wild type EP24.15 has a k(cat)/K(M) of 2.5x10(4) M(-1) s(-1) for this substrate.  相似文献   

16.
The solvent kinetic isotope effects (SKIE) on the yeast alpha-glucosidase-catalyzed hydrolysis of p-nitrophenyl and methyl-d-glucopyranoside were measured at 25 degrees C. With p-nitrophenyl-D-glucopyranoside (pNPG), the dependence of k(cat)/K(m) on pH (pD) revealed an unusually large (for glycohydrolases) solvent isotope effect on the pL-independent second-order rate constant, (DOD)(k(cat)/K(m)), of 1.9 (+/-0.3). The two pK(a)s characterizing the pH profile were increased in D(2)O. The shift in pK(a2) of 0.6 units is typical of acids of comparable acidity (pK(a)=6.5), but the increase in pK(a1) (=5.7) of 0.1 unit in going from H(2)O to D(2)O is unusually small. The initial velocities show substrate inhibition (K(is)/K(m) approximately 200) with a small solvent isotope effect on the inhibition constant [(DOD)K(is)=1.1 (+/-0.2)]. The solvent equilibrium isotope effects on the K(is) for the competitive inhibitors D-glucose and alpha-methyl D-glucoside are somewhat higher [(DOD)K(i)=1.5 (+/-0.1)]. Methyl glucoside is much less reactive than pNPG, with k(cat) 230 times lower and k(cat)/K(m) 5 x 10(4) times lower. The solvent isotope effect on k(cat) for this substrate [=1.11 (+/-0. 02)] is lower than that for pNPG [=1.67 (+/-0.07)], consistent with more extensive proton transfer in the transition state for the deglucosylation step than for the glucosylation step.  相似文献   

17.
Salts inhibit the activity of sweet almond beta-glucosidase. For cations (Cl(-) salts) the effectiveness follows the series: Cu(+2), Fe(+2)>Zn(+2)>Li(+)>Ca(+2)>Mg(+2)>Cs(+)>NH(4)(+)>Rb(+)>K(+)>Na(+) and for anions (Na(+) salts) the series is: I(-)>ClO(4)(-)>(-)SCN>Br(-) approximately NO(3)(-)>Cl(-) approximately (-)OAc>F(-) approximately SO(4)(-2). The activity of the enzyme, like that of most glycohydrolases, depends on a deprotonated carboxylate (nucleophile) and a protonated carboxylic acid for optimal activity. The resulting pH-profile of k(cat)/K(m) for the beta-glucosidase-catalyzed hydrolysis of p-nitrophenyl glucoside is characterized by a width at half height that is strongly sensitive to the nature and concentration of the salt. Most of the inhibition is due to a shift in the enzymic pK(a)s and not to an effect on the pH-independent second-order rate constant, (k(cat)/K(m))(lim). For example, as the NaCl concentration is increased from 0.01 M to 1.0 M the apparent pK(a1)increases (from 3.7 to 4.9) and the apparent pK(a2)decreases (from 7.2 to 5.9). With p-nitrophenyl glucoside, the value of the pH-independent (k(cat)/K(m))(lim) (=9 x 10(4) M(-1) s(-1)) is reduced by less than 4% as the NaCl concentration is increased. There is a similar shift in the pK(a)s when the LiCl concentration is increased to 1.0 M. The results of these salt-induced pK(a) shifts rule out a significant contribution of reverse protonation to the catalytic efficiency of the enzyme. At low salt concentration, the fraction of the catalytically active monoprotonated enzyme in the reverse protonated form (i.e., proton on the group with a pK(a) of 3.7 and dissociated from the group with a pK(a) of 7.2) is very small ( approximately 0.03%). At higher salt concentrations, where the two pK(a)s become closer, the fraction of the monoprotonated enzyme in the reverse protonated form increases over 300-fold. However, there is no increase in the intrinsic reactivity, (k(cat)/K(m))(lim), of the monoprotonated species. For other enzymes which may show such salt-induced pK(a) shifts, this provides a convenient test for the role of reverse protonation.  相似文献   

18.
Ruzzini AC  Horsman GP  Eltis LD 《Biochemistry》2012,51(29):5831-5840
meta-Cleavage product (MCP) hydrolases catalyze C-C bond fission in the aerobic catabolism of aromatic compounds by bacteria. These enzymes utilize a Ser-His-Asp triad to catalyze hydrolysis via an acyl-enzyme intermediate. BphD, which catalyzes the hydrolysis of 2-hydroxy-6-oxo-6-phenylhexa-2,4-dienoic acid (HOPDA) in biphenyl degradation, catalyzed the hydrolysis of an ester analogue, p-nitrophenyl benzoate (pNPB), with a k(cat) value (6.3 ± 0.5 s(-1)) similar to that of HOPDA (6.5 ± 0.5 s(-1)). Consistent with the breakdown of a shared intermediate, product analyses revealed that BphD catalyzed the methanolysis of both HOPDA and pNPB, partitioning the products to benzoic acid and methyl benzoate in similar ratios. Turnover of HOPDA was accelerated up to 4-fold in the presence of short, primary alcohols (methanol > ethanol > n-propanol), suggesting that deacylation is rate-limiting during catalysis. In the steady-state hydrolysis of HOPDA, k(cat)/K(m) values were independent of methanol concentration, while both k(cat) and K(m) values increased with methanol concentration. This result was consistent with a simple model of nucleophilic catalysis. Although the enzyme could not be saturated with pNPB at methanol concentrations of >250 mM, k(obs) values from the steady-state turnover of pNPB at low methanol concentrations were also consistent with a nucleophilic mechanism of catalysis. Finally, transient-state kinetic analysis of pNPB hydrolysis by BphD variants established that substitution of the catalytic His reduced the rate of acylation by more than 3 orders of magnitude. This suggests that for pNPB hydrolysis, the serine nucleophile is activated by the His-Asp dyad. In contrast, rapid acylation of the H265Q variant during C-C bond cleavage suggests that the serinate forms via a substrate-assisted mechanism. Overall, the data indicate that ester hydrolysis proceeds via the same acyl-enzyme intermediate as that of the physiological substrate but that the serine nucleophile is activated via a different mechanism.  相似文献   

19.
The broadly reactive cysteine protease dipeptidyl peptidase I (DPPI, cathepsin C) is thought to activate all progranzymes (zymogens of lymphocyte serine proteases) to form mature granzymes. We synthesized dipeptide 7-amino-4-methylcoumarin (AMC) substrates containing progranzyme activation sequences and showed that they were efficiently hydrolyzed by DPPI. However, DPPI will not hydrolyze Ile-Ile-AMC, the N-terminal dipeptide sequence found in mature granzymes. Introduction of the nonphysiological homophenylalanine (Hph) residue at P1 resulted in the best substrate Ala-Hph-AMC for DPPI (k(cat)/K(m)=9,000,000M(-1)s(-1)). The charged N-terminal amino group of the substrate was essential and replacement of the NH(2) group with OH or NH(CH(3)) in Gly-Phe-AMC reduced the k(cat)/K(m) value by two to three orders of magnitude. A hydrazide azaglycine analog, NH(2)NHCO-Phe-AMC, was not hydrolyzed at pH 5.5, but underwent slow hydrolysis at lower pHs where the amino group is partially protonated. DPPI also failed to hydrolyze NH(2)COCH(2)-Phe-AMC, where the NH(2) group is unprotonated. The results reported in this paper should be useful in the design of better DPPI inhibitors to block granzyme maturation and granzyme-dependent apoptosis.  相似文献   

20.
Bile salt-dependent lipase was purified to homogeneity from lyophilized human milk and used to screen the influence of the acyl chain length (2-16 carbon atoms) on the kinetic constants k(cat) and K(m) of the hydrolysis of para-nitrophenyl (pnp) ester substrates in the presence or absence of sodium taurocholate (NaTC: 0.02-20 mM). The highest k(cat) value (~3,500 s(-1)) was obtained with pnpC(8) as substrate, whereas the lowest K(m) (<10 μM) was that recorded with pnpC(10). In the absence of NaTC, the maximal catalytic efficiency (k(cat)/K(m)) was obtained with pnpC(8), while in the presence of NaTC k(cat)/K(m) was maximal with pnpC(8), pnpC(10) or pnpC(12). The bile salt activated the enzyme in two successive saturation phases occurring at a micromolar and a millimolar concentration range, respectively. The present data emphasize the suitability of this enzyme for the hydrolysis of medium-chain acyl-containing substrates and throw additional light on how BSDL is activated by NaTC.  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