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.
To elucidate a detailed catalytic mechanism for nitrile hydratases (NHases), the pH and temperature dependence of the kinetic constants k(cat) and K(m) for the cobalt-type NHase from Pseudonocardia thermophila JCM 3095 (PtNHase) were examined. PtNHase was found to exhibit a bell-shaped curve for plots of relative activity versus pH at pH 3.2-11 and was found to display maximal activity between pH 7.2 and 7.8. Fits of these data provided pK(E)(S1) and pK(E)(S2) values of 5.9 +/- 0.1 and 9.2 +/- 0.1 (k(cat)' = 130 +/- 1 s(-1)), respectively, and pK(E)(1) and pK(E)(2) values of 5.8 +/- 0.1 and 9.1 +/- 0.1 (k(cat)'/K(m)' = (6.5 +/- 0.1) x 10(3) s(-1) mm(-1)), respectively. Proton inventory studies indicated that two protons are transferred in the rate-limiting step of the reaction at pH 7.6. Because PtNHase is stable at 60 degrees C, an Arrhenius plot was constructed by plotting ln(k(cat)) versus 1/T, providing E(a) = 23.0 +/- 1.2 kJ/mol. The thermal stability of PtNHase also allowed DeltaH(0) ionization values to be determined, thus helping to identify the ionizing groups exhibiting the pK(E)(S1) and pK(E)(S2) values. Based on DeltaH(0)(ion) data, pK(E)(S1) is assigned to betaTyr(68), whereas pK(E)(S2) is assigned to betaArg(52), betaArg(157), or alphaSer(112) (NHases are alpha(2)beta(2)-heterotetramers). A combination of these data with those previously reported for NHases and synthetic model complexes, along with sequence comparisons of both iron- and cobalt-type NHases, allowed a novel catalytic mechanism for NHases to be proposed.  相似文献   

2.
1. N-Benzoyl-l-serine methyl ester was synthesized and evaluated as a substrate for bromelain (EC 3.4.22.4) and for papain (EC 3.4.22.2). 2. For the bromelain-catalysed hydrolysis at pH7.0, plots of [S(0)]/v(i) (initial substrate concn./initial velocity) versus [S(0)] are markedly curved, concave downwards. 3. Analysis by lattice nomography of a modifier kinetic mechanism in which the modifier is substrate reveals that concave-down [S(0)]/v(i) versus [S(0)] plots can arise when the ratio of the rate constants that characterize the breakdown of the binary (ES) and ternary (SES) complexes is either less than or greater than 1. In the latter case, there are severe restrictions on the values that may be taken by the ratio of the dissociation constants of the productive and non-productive binary complexes. 4. Concave-down [S(0)]/v(i) versus [S(0)] plots cannot arise from compulsory substrate activation. 5. Computational methods, based on function minimization, for determination of the apparent parameters that characterize a non-compulsory substrate-activated catalysis are described. 6. In an attempt to interpret the catalysis by bromelain of the hydrolysis of N-benzoyl-l-serine methyl ester in terms of substrate activation, the general substrate-activation model was simplified to one in which only one binary ES complex (that which gives rise directly to products) can form. 7. In terms of this model, the bromelain-catalysed hydrolysis of N-benzoyl-l-serine methyl ester at pH7.0, I=0.1 and 25 degrees C is characterized by K(m) (1) (the dissociation constant of ES)=1.22+/-0.73mm, k (the rate constant for the breakdown of ES to E+products, P)=1.57x10(-2)+/-0.32x10(-2)s(-1), K(a) (2) (the dissociation constant that characterizes the breakdown of SES to ES and S)=0.38+/-0.06m, and k' (the rate constant for the breakdown of SES to E+P+S)=0.45+/-0.04s(-1). 8. These parameters are compared with those in the literature that characterize the bromelain-catalysed hydrolysis of alpha-N-benzoyl-l-arginine ethyl ester and of alpha-N-benzoyl-l-arginine amide; K(m) (1) and k for the serine ester hydrolysis are somewhat similar to K(m) and k(cat.) for the arginine amide hydrolysis and K(as) and k' for the serine ester hydrolysis are somewhat similar to K(m) and k(cat.) for the arginine ester hydrolysis. 9. A previous interpretation of the inter-relationships of the values of k(cat.) and K(m) for the bromelain-catalysed hydrolysis of the arginine ester and amide substrates is discussed critically and an alternative interpretation involving substantial non-productive binding of the arginine amide substrate to bromelain is suggested. 10. The parameters for the bromelain-catalysed hydrolysis of the serine ester substrate are tentatively interpreted in terms of non-productive binding in the binary complex and a decrease of this type of binding by ternary complex-formation. 11. The Michaelis parameters for the papain-catalysed hydrolysis of the serine ester substrate (K(m)=52+/-4mm, k(cat.)=2.80+/-0.1s(-1) at pH7.0, I=0.1, 25.0 degrees C) are similar to those for the papain-catalysed hydrolysis of methyl hippurate. 12. Urea and guanidine hydrochloride at concentrations of 1m have only small effects on the kinetic parameters for the hydrolysis of the serine ester substrate catalysed by bromelain and by papain.  相似文献   

3.
4.
Badarau A  Page MI 《Biochemistry》2006,45(35):10654-10666
The kinetics and mechanism of hydrolysis of the native zinc and metal substituted Bacillus cereus (BcII) metallo-beta-lactamase have been investigated. The pH and metal ion dependence of k(cat) and k(cat)/K(m), determined under steady-state conditions, for the cobalt substituted BcII catalyzed hydrolysis of cefoxitin, cephaloridine, and cephalexin indicate that an enzyme residue of apparent pK(a) 6.3 +/- 0.1 is required in its deprotonated form for metal ion binding and catalysis. The k(cat)/K(m) for cefoxitin and cephalexin with cadmium substituted BcII is dependent on two ionizing groups on the enzyme: one of pK(a1) = 8.7 +/- 0.1 required in its deprotonated form and the other of pK(a2) = 9.3 +/- 0.1 required in its protonated form for activity. The pH dependence of the competitive inhibition constant, K(i), for CdBcII with l-captopril indicates that pK(a1) = 8.7 +/- 0.1 corresponds to the cadmium-bound water. For the manganese substituted BcII, the pH dependence of k(cat)/K(m) for benzylpenicillin, cephalexin, and cefoxitin similarly indicated the importance of two catalytic groups: one of pK(a1) = 8.5 +/- 0.1 which needs to be deprotonated and the other of pK(a2) = 9.4 +/- 0.1 which needs to be protonated for catalysis; the pK(a1) was assigned to the manganese-bound water. The rate was metal ion concentration dependent at the highest manganese concentrations used (10(-)(3) M). The metal substituted species have similar or higher catalytic activities compared with the zinc enzyme, albeit at pHs above 7. Interestingly, with cefoxitin, a very poor substrate for ZnBcII, both k(cat) and k(cat)/K(m) increase with increasing pK(a) of the metal-bound water, in the order Zn < Co < Mn < Cd. A higher pK(a) for the metal-bound water for cadmium and manganese BCII leads to more reactive enzymes than the native zinc BcII, suggesting that the role of the metal ion is predominantly to provide the nucleophilic hydroxide, rather than to act as a Lewis acid to polarize the carbonyl group and stabilize the oxyanion tetrahedral intermediate.  相似文献   

5.
Hardy LW  Nishida CH  Kirsch JF 《Biochemistry》1984,23(6):1288-1294
The pH dependence of k(cat) for the Bacillus cereus beta-lactamase I catalyzed hydrolysis of carbenicillin(VI), which differs from benzylpenicillin (I) in having a carboxylic moiety alpha to the phenyl ring, exhibits a profile consistent with a model in which the alpha-COOH and alpha-COO forms of the ES complex turn over with respective rate constants of 2152 s(-1) and 384 s(-1). The pK(a)(app) for the alpha-COOH is shifted from 3.2 in solution to 6.1 in the ES complex. The normalized k(cat)/K(m) vs. pH profile for VI is not superimposable on that of I, indicating that both the neutral and anionic forms of the carboxyl moiety of VI combine with the enzyme to give the first irreversibly formed complex, presumably the acyl-enzyme. Quantitative accord with the kinetic data is achieved only through fitting to a model where kinetically significant proton transfer in the ES complex is permitted. The second-order rate constants for the reaction of the enzyme with the alpha-COOH and alpha-COO forms of VI are 2.2 x 10(8) M(-1) s(-1) and 3.8 x 10(6) M(-1) s(-1), respectively. The high value for the alpha-COOH form suggests that this reaction may be in part diffusion controlled. This conjecture is borne out by the observation that the sensitivity of k(cat)/K(m) to eta(rel) decreases with increasing pH for VI, whereas this sensitivity is pH independent for I. These conclusions are further supported by the results of a kinetic investigation of the pH dependence of sulbenicillin (VII) where an alpha-SO3H replaces the alpha-COOH of VI. The strongly acidic sulfonic acid moiety of VII is fully ionized throughout nearly the entire pH range of interest, and its kinetics, as a function of pH, are very similar to those observed and calculated for the alpha-COO form of VI. Solvent deuterium kinetic isotope effects are reported for k(cat) and k(cat)/K(m) for both VI and VII.  相似文献   

6.
Wang J  Edmondson DE 《Biochemistry》2011,50(35):7710-7717
Monoamine oxidase A (MAO A) is a mitochondrial outer membrane-bound flavoenzyme important in the regulation of serotonin and dopamine levels. Because the rat is extensively used as an animal model in drug studies, it is important to understand how rat MAO A behaves in comparison with the more extensively studied human enzyme. For many reversible inhibitors, rat MAO A exhibits K(i) values similar to those of human MAO A. The pH profile of k(cat) for rat MAO A shows a pK(a) of 8.2 ± 0.1 for the benzylamine ES complex and pK(a) values of 7.5 ± 0.1 and 7.6 ± 0.1 for the ES complexes with p-CF(3)-(1)H- and p-CF(3)-(2)H-benzylamine, respectively. In contrast to the human enzyme, the rat enzyme exhibits a single pK(a) value (8.3 ± 0.1) with k(cat)/K(m) for benzylamine versus pH and pK(a) values of 7.8 ± 0.1 and 8.1 ± 0.2 for the ascending limbs, respectively, of k(cat)/K(m) versus pH profiles for p-CF(3)-(1)H- and p-CF(3)-(2)H-benzylamine and 9.3 ± 0.1 and 9.1 ± 0.2 for the descending limbs, respectively. The oxidation of para-substituted benzylamine substrate analogues by rat MAO A has large deuterium kinetic isotope effects on k(cat) and on k(cat)/K(m). These effects are pH-independent and range from 7 to 14, demonstrating a rate-limiting α-C-H bond cleavage step in catalysis. Quantitative structure-activity correlations of log k(cat) with the electronic substituent parameter (σ) at pH 7.5 and 9.0 show a dominant contribution with positive ρ values (1.2-1.3) and a pH-independent negative contribution from the steric term. Quantitative structure-activity relationship analysis of the binding affinities of the para-substituted benzylamine analogues for rat MAO A shows an increased van der Waals volume (V(w)) increases the affinity of the deprotonated amine for the enzyme. These results demonstrate that rat MAO A exhibits functional properties similar but not identical with those of the human enzyme and provide additional support for C-H bond cleavage via a polar nucleophilic mechanism.  相似文献   

7.
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   

8.
Nitrile hydratase (NHase) has attracted substantial attention for industrial applications to produce large-scale amides. Several NHases have been investigated for functional expression in Escherichia coli (E. coli). A Fe-type NHase was obtained from an acetamiprid-degrading bacterium, Pseudoxanthomonas sp. AAP-7 and functionally expressed in E. coli BL21 (DE3). No significant NHase activity was detected from the E. coli expressing either the NHase gene alone or NHase and P46K genes transcribed as one unit. Purified recombinant NHase, co-expressed with P46K on two separate plasmids, exhibited the maximal enzyme activity. Furthermore, a GST tag attached to the N-terminus of α subunit resulted in a slight increase in the solubility and stability of NHase compared with a His tag at the C-terminus of β subunit. When co-expressed with the chaperones GroEL-GroES, the yield of the soluble recombinant NHase was improved substantially, while a small decrease in NHase activity was observed. The putative activator P46K was strictly required for production of the recombinant NHase for full enzyme activity, although the chaperones GroEL-GroES appeared to assist NHase to fold properly. This study of the expression of a fully active Fe-type NHase would provide another example to enhance our understanding of NHase biosynthesis.  相似文献   

9.
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   

10.
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   

11.
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   

12.
The PutA flavoprotein from Escherichia coli combines DNA-binding, proline dehydrogenase (PRODH), and Delta(1)-pyrroline-5-carboxylate dehydrogenase (P5CDH) activities onto a single polypeptide. Recently, an X-ray crystal structure of PutA residues 87-612 was solved which identified a D370-Y540 hydrogen bond pair in the PRODH active site that appears to have an important role in shaping proline binding and the FAD redox environment. To examine the role of D370-Y540 in the PRODH active site, mutants D370A, Y540F, and D370A/Y540F were characterized in a form of PutA containing only residues 86-601 (PutA86-601) designed to mimic the known structural region of PutA (87-612). Disruption of the D370-Y540 pair only slightly diminished k(cat), while more noticeable affects were observed in K(m). The mutant D370A/Y540F showed the most significant changes in the pH dependence of k(cat)/K(m) and K(m) relative to wild-type PutA86-601 with an apparent pK(a) value of about 8.2 for the pH-dependent decrease in K(m). From the pH profile of D370A/Y540F inhibition by l-tetrahydro-2-furoic acid (l-THFA), the pH dependency of K(m) in D370A/Y540F is interpreted as resulting from the deprotonation of the proline amine in the E-S complex. Replacement of D370 and Y540 produces divergent effects on the E(m) for bound FAD. At pH 7.0, E(m) values of -0.026, -0.089 and -0.042 V were determined for the two-electron reduction of bound FAD in D370A, Y540F and D370A/Y540F, respectively. The 40-mV positive shift in E(m) determined for D370A relative to wild-type PutA86-601 (E(m)=-0.066 V, pH 7.0) indicates D370 has a key role in modulating the FAD redox environment.  相似文献   

13.
Recombinant LytB protein from the thermophilic eubacterium Aquifex aeolicus produced in Escherichia coli was purified to apparent homogeneity. The purified LytB protein catalyzed the reduction of (E)-4-hydroxy-3-methyl-but-2-enyl diphosphate (HMBPP) in a defined in vitro system. The reaction products were identified as isopentenyl diphosphate and dimethylallyl diphosphate. A spectrophotometric assay was established to determine the steady-state kinetic parameters of LytB protein. The maximal specific activity of 6.6+/-0.3 micromol x min(-1) x mg(-1) protein was determined at pH 7.5 and 60 degrees C. The k(cat) value of the LytB protein was 3.7+/-0.2 s(-1) and the K(m) value for HMBPP was 590+/-60 microM.  相似文献   

14.
Hydrolytic reactions of oligopeptide 4-nitroanilides catalyzed by human-alpha-thrombin, human activated protein C and human factor Xa were studied at pH 8.0-8.4 and 25.0+/-0.1 degrees C by the progress curve method and individual rate constants were calculated mostly within 10% internal error using DYNAFITV. A systematic strategy has been developed for fitting a three-step consecutive mechanism to eighteen hundred to six thousand time-course data points polled from two to four independent kinetic experiments. Enzyme and substrate concentrations were also calculated. Individual rate constants well reproduce published values obtained under comparable conditions and the Michaelis-Menten kinetic parameters calculated from these elementary rate constants are also within reasonable limits of published values. For comparison, the integrated Michaelis-Menten equation was also fitted to data from twelve sets. Both the k(cat) and k(cat)/K(m) values are within 15% agreement with those calculated using the elementary rate constants obtained with DYNAFITV. Rate constants for the second and third consecutive steps are within 3-4 fold indicating that both determine the overall rate. The Factor Xa-catalyzed hydrolysis of N-alpha-Z-D-Arg-Gly-Arg-pNA.2HCl at pH 8.4 in a series of buffers containing increasing fractions of deuterium at 25.0+/-0.1 degrees C shows a very strong dependence of k(3) and a moderate dependence of k(2) on D content in the buffer: the fractionation factors are: 0.49+/-0.03 for K(1,) 0.70+/-0.05 for k(2), and (0.32+/-0.03)(2) for k(3).  相似文献   

15.
We report herein the functional expression of an Fe-type nitrile hydratase (NHase) without the co-expression of an activator protein or the Escherichia coli chaperone proteins GroES/EL. Soluble protein was obtained when the α- and β-subunit genes of the Fe-type NHase Comamonas testosteroni Ni1 (CtNHase) were synthesized with optimized E. coli codon usage and co-expressed. As a control, the Fe-type NHase from Rhodococcus equi TG328–2 (ReNHase) was expressed with (ReNHase+Act) and without (ReNHase?Act) its activator protein, establishing that expression of a fully functional, metallated ReNHase enzyme requires the co-expression of its activator protein, similar to all other Fe-type NHase enzymes reported to date, whereas the CtNHase does not. The X-ray crystal structure of CtNHase was determined to 2.4 Å resolution revealing an αβ heterodimer, similar to other Fe-type NHase enzymes, except for two important differences. First, two His residues reside in the CtNHase active site that are not observed in other Fe-type NHase enzymes and second, the active site Fe(III) ion resides at the bottom of a wide solvent exposed channel. The solvent exposed active site, along with the two active site histidine residues, are hypothesized to play a role in iron incorporation in the absence of an activator protein.  相似文献   

16.
Hardy LW  Kirsch JF 《Biochemistry》1984,23(6):1282-1287
The solvent kinetic isotope effects (SKIE's) on k(cat) (D(V)) and on k(cat/Km[D(V/K)] were determined for the Bacillus cereus beta-lactamase I catalyzed hydrolysis of five substrates that have values of k(cat)/K(m) varying over the range (0.014-46.3) X 10(6)M(-1) s(-1) and of k(cat) between 0.5 and 2019 s(-1). The variation of D(V/K) was only from 1.06 to 1.25 among these compounds and that in D(V) was from 1.50 to 2.16. These results require that Dk(1), the SKIE on the enzyme-substrate association rate constant, and D(k-1/k2), that on the partition ratio of the ES complex, both be near 1. The larger SKIE observed on D(V) requires that an exchangeable proton be in flight for either or both the acylation and the deacylation reaction. The pH dependence of the values k(cat)/K(m) for three substrates shows identical pK(a)s of 5.5. and 8.4. This identity combined with the fact that only one of these three substrates is kinetically "sticky" proves that the substrates can combine productively with only one protonic form of the enzyme. There is considerable substrate variation in the pK(a) values of k(cat) observed vs. pH profiles; the inflection points for all substrates studied are at pH values more extreme than are observed in the pH profiles for k(cat)/K(m).  相似文献   

17.
The first committed step of lipid A biosynthesis in Gram-negative bacteria is catalyzed by the zinc-dependent hydrolase LpxC that removes an acetate from the nitrogen at the 2' '-position of UDP-3-O-acyl-N-acetylglucosamine. Recent structural characterization by both NMR and X-ray crystallography provides many important details about the active site environment of LpxC from Aquifex aeolicus, a heat-stable orthologue that displays 32% sequence identity to LpxC from Escherichia coli. The detailed reaction mechanism and specific roles of active site residues for LpxC from A. aeolicus are further analyzed here. The pH dependencies of k(cat)/K(M) and k(cat) for the deacetylation of the substrate UDP-3-O-[(R)-3-hydroxymyristoyl]-GlcNAc are both bell-shaped. The ascending acidic limb (pK(1)) was fitted to 6.1 +/- 0.2 for k(cat) and 5.7 +/- 0.2 for k(cat)/K(M). The descending basic limb (pK(2)) was fitted to 8.0 +/- 0.2 for k(cat) and 8.4 +/- 0.2 for k(cat)/K(M). The pH dependence of the E73A mutant exhibits loss of the acidic limb, and the mutant retains only 0.15% activity versus the wild type. The pH dependencies of the other active site mutants H253A, K227A, H253A/K227A, and D234N remain bell-shaped, although their significantly lower activities (0.25%, 0.05%, 0.007%, and 0.57%, respectively) suggest that they contribute significantly to catalysis. Our cumulative data support a mechanism for LpxC wherein Glu73 serves as the general base for deprotonation and activation of the zinc-bound water.  相似文献   

18.
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   

19.
Yang Z  Lu CD 《Journal of bacteriology》2007,189(11):3954-3959
The arginine transaminase (ATA) pathway represents one of the multiple pathways for L-arginine catabolism in Pseudomonas aeruginosa. The AruH protein was proposed to catalyze the first step in the ATA pathway, converting the substrates L-arginine and pyruvate into 2-ketoarginine and L-alanine. Here we report the initial biochemical characterization of this enzyme. The aruH gene was overexpressed in Escherichia coli, and its product was purified to homogeneity. High-performance liquid chromatography and mass spectrometry (MS) analyses were employed to detect the presence of the transamination products 2-ketoarginine and L-alanine, thus demonstrating the proposed biochemical reaction catalyzed by AruH. The enzymatic properties and kinetic parameters of dimeric recombinant AruH were determined by a coupled reaction with NAD(+) and L-alanine dehydrogenase. The optimal activity of AruH was found at pH 9.0, and it has a novel substrate specificity with an order of preference of Arg > Lys > Met > Leu > Orn > Gln. With L-arginine and pyruvate as the substrates, Lineweaver-Burk plots of the data revealed a series of parallel lines characteristic of a ping-pong kinetic mechanism with calculated V(max) and k(cat) values of 54.6 +/- 2.5 micrromol/min/mg and 38.6 +/- 1.8 s(-1). The apparent K(m) and catalytic efficiency (k(cat)/K(m)) were 1.6 +/- 0.1 mM and 24.1 mM(-1) s(-1) for pyruvate and 13.9 +/- 0.8 mM and 2.8 mM(-1) s(-1) for l-arginine. When L-lysine was used as the substrate, MS analysis suggested Delta(1)-piperideine-2-carboxylate as its transamination product. These results implied that AruH may have a broader physiological function in amino acid catabolism.  相似文献   

20.
Streptomyces caespitosus neutral protease (ScNP) is one of the smallest metalloproteinase with a molecular mass of 14 kDa. Effects of solvent composition on ScNP activity were examined using a peptide substrate. The k(cat)/K(m) values of ScNP exhibited bell-shaped pH-dependence with the optimal pH of 6.4-7.0 and the pK(a) values of 5.0 +/- 0.1 and 8.3 +/- 0.1. ScNP activity increased in an exponential fashion with increasing [NaCl]. The relative k(cat)/K(m) value at 3.6 M NaCl to that at 0 M NaCl was 3.7, and the degree of the activation at x M NaCl was expressed as 1.2 (x) (x < 2.0) and 1.4(x) (x > 2.0). On the other hand, ScNP activity decreased with increasing concentrations of LiCl, KCl, NaBr, LiBr, KBr and NaClO(4). Alcohols inhibited ScNP activity with the IC(50) values, the concentration required for decreasing the activity at 50% of the maximum, of 0.77-6.54 M. The order of the inhibitory potency was 1-butanol, 2-methyl-1-propanol, 2-methyl-2-butanol > 2-methyl-2-propanol, 2-butanol, 1-propanol > 2-propanol > ethanol > methanol. The activities recovered completely by the dilution of alcohols, suggesting that the ScNP inhibition by alcohols is reversible. These characteristics of ScNP are compared with those of human matrix metalloproteinase 7 and thermolysin.  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