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 328 毫秒
1.
The gene encoding alanine dehydrogenase (AlaDH; EC 1.4.1.1) from the marine psychrophilic bacterium strain PA-43 was cloned, sequenced, and overexpressed in Escherichia coli. The primary structure was deduced on the basis of the nucleotide sequence. The enzyme subunit contains 371 amino acid residues, and the sequence is 90% and 77% identical, respectively, to AlaDHs from Shewanella Ac10 and Vibrio proteolyticus. The half-life of PA-43 AlaDH at 52 degrees C is 9 min, and it is thus more thermolabile than the AlaDH from Shewanella Ac10 or V. proteolyticus. The enzyme showed strong specificity for NAD(+) and l-alanine as substrates. The apparent K(m) for NAD(+) was temperature dependent (0.04 mM-0.23 mM from 15 degrees C to 55 degrees C). A comparison of the PA-43 deduced amino acid sequence to the solved three-dimensional structure of Phormidium lapideum AlaDH showed that there were likely to be fewer salt bridges in the PA-43 enzyme, which would increase enzyme flexibility and decrease thermostability. The hydrophobic surface character of the PA-43 enzyme was greater than that of P. lapideum AlaDH, by six residues. However, no particular modification or suite of modifications emerged as being clearly responsible for the psychrophilic character of PA-43 AlaDH.  相似文献   

2.
A novel protease designated protease-A-17N-1, was purified from the halo-alkalophilic Bacillus sp. 17N-1, and found active in media containing dithiothreitol and EDTAK(2). This enzyme maintained significant activity from pH 6.00 to 9.00, showed optimum k(cat)/K(m) value at pH 7.50 and 33 degrees C. It was observed that only specific inhibitors of cysteine proteinases inhibited its activity. The pH-(k(cat)/K(m)) profile of protease-A-17N-1 was described by three pK(a)s in the acid limb, and one in the alkaline limb. Both are more likely due t3o the protonic dissociation of an acidic residue, and the development and subsequent deprotonation of an ion-pair, respectively, in its catalytic site, characteristic for cysteine proteinases. Moreover, both the obtained estimates of rate constant k(1) and the ratio k(2)/k(-1) at 25 degrees C, from the temperature-(k(cat)/K(m)) profile of protease-A-17N-1, were found similar to those estimated from the proton inventories of the same parameter, verifying the reliability of the latter methodology. Besides, the bowed-downward proton inventories of k(cat)/K(m), as well as the large inverse SIE observed for this parameter, in combination with its dependence versus temperature, were showed unambiguously that k(cat)/K(m) = k(1). Such results suggest that the novel enzyme is more likely to be a cysteine proteinase functioning via a general acid-base mechanism.  相似文献   

3.
Mouse thymidylate synthase R209K (a mutation corresponding to R218K in Lactobacillus casei), overexpressed in thymidylate synthase-deficient Escherichia coli strain, was poorly soluble and with only feeble enzyme activity. The mutated protein, incubated with FdUMP and N(5,10)-methylenetetrahydrofolate, did not form a complex stable under conditions of SDS/polyacrylamide gel electrophoresis. The reaction catalyzed by the R209K enzyme (studied in a crude extract), compared to that catalyzed by purified wild-type recombinant mouse thymidylate synthase, showed the K(m) value for dUMP 571-fold higher and V(max) value over 50-fold (assuming that the mutated enzyme constituted 20% of total crude extract protein) lower. Thus the ratios k(cat, R209K)/k(cat, 'wild') and (k(cat, R209K)/K(m, R209K)(dUMP))/( k(cat, 'wild')/K(m, 'wild')(dUMP)) were 0.019 and 0.000032, respectively, documenting that mouse thymidylate synthase R209, similar to the corresponding L. casei R218, is essential for both dUMP binding and enzyme reaction.  相似文献   

4.
Alanine dehydrogenase (AlaDH: EC 1.4.1.1), malate dehydrogenase (MDH: EC 1.1.1.37), and glutamate dehydrogenase (EC 1.4.1.2), all NAD+ dependent, were detected in extracts from a psychrophilic bacterium, strain PA-43, isolated from a sea urchin off the Icelandic coast. Characterization tests suggested that the strain had a close relationship to Vibrio, but sequencing of part of the 16S rDNA gene placed the bacterium among Shewanella species in a constructed phylogenetic tree. The bacterium had an optimum growth temperature of 16.5 degrees C, and maximum dehydrogenase expression was obtained in a rich medium supplemented with NaCl. Both AlaDH and MDH were purified to homogeneity. AlaDH is a hexamer, with an approximate relative molecular mass of 260,000, whereas MDH is dimeric, with an apparent relative molecular mass of approximately 70,000. Both enzymes were thermolabile, and the optimum temperatures for activity were shifted toward lower temperatures than those found in the same enzymes from mesophiles, 37 degrees C for MDH and approximately 47 degrees C for AlaDH. The pH optima for AlaDH in the forward and reverse reactions were 10.5 and 9, respectively, whereas those for MDH were 10-10.2 and 8.8, respectively. Partial amino acid sequences, comprising approximately 30% of the total sequences from each enzyme, were determined for N-terminal, tryptic, and chymotryptic fragments of the enzymes. The AlaDH showed the highest similarity to AlaDHs from the psychrotroph Shewanella Ac10 and the mesophile Vibrio proteolyticus, whereas MDH was most similar to the MDHs from the mesophiles Escherichia coli and Haemophilus influenzae, with lower identity to the psychrophilic malate dehydrogenases from Vibrio 5710 and Photobacterium SS9.  相似文献   

5.
A trehalose-6-phosphate phosphatase (TPP) gene, otsB, from a psychrotrophic bacterium, Arthrobacter strain A3, was identified. The product of this otsB gene is 266 amino acids in length with a calculated molecular weight of 27,873 Da. The protein was expressed in Escherichia coli and purified to apparent homogeneity. The purified recombinant TPP catalyzed the dephosphorylation of trehalose-6-phosphate to form trehalose and showed a broad optimum pH range from 5.0 to 7.5. This enzyme also showed an absolute requirement for Mg(2+) or Co(2+) for catalytic activity. The recombinant TPP had a maximum activity at 30 °C and maintained activity over a temperature range of 4-30 °C. TPP was generally heat-labile, losing 70 % of its activity when subjected to heat treatment at 50 °C for 6 min. Kinetic analysis of the Arthrobacter strain A3 TPP showed ~tenfold lower K (m) values when compared with values derived from other bacterial TPP enzymes. The highest k (cat)/K (m) value was 37.5 mM(-1) s(-1) (repeated three times), which is much higher than values published for mesophilic E. coli TPP, indicating that the Arthrobacter strain A3 TPP possessed excellent catalytic activity at low temperatures. Accordingly, these characteristics suggest that the TPP from the Arthrobacter strain A3 is a new cold-adapted enzyme. In addition, this is the first report characterizing the enzymatic properties of a TPP from a psychrotrophic organism.  相似文献   

6.
A psychrotrophic bacterium Shewanella sp. strain SIB1 was grown at 4 and 20 degrees C, and total soluble proteins extracted from the cells were analyzed by two-dimensional polyacrylamide gel electrophoresis. Comparison of these patterns showed that the cellular content of a protein with a molecular mass of 28 kDa and an isoelectric point of four greatly increased at 4 degrees C compared to that at 20 degrees C. Determination of the N-terminal amino acid sequence, followed by the cloning and sequencing of the gene encoding this protein, revealed that this protein is a member of the FKBP family of proteins with an amino acid sequence identity of 56% to Escherichia coli FKBP22. This protein was overproduced in E. coli in a His-tagged form, purified, and analyzed for peptidyl-prolyl cis-trans isomerase activity. When this activity was determined by the protease coupling assay using N-succinyl-Ala-Leu-Pro-Phe-p-nitroanilide as a substrate at various temperatures, the protein exhibited the highest activity at 10 degrees C with a k(cat)/K(m) value of 0.87 micro m(-1) x s(-1). When the peptidyl-prolyl cis-trans isomerase activity was determined by the RNase T(1) refolding assay at 10 and 20 degrees C, the protein exhibited higher activity at 10 degrees C with a k(cat)/K(m) value of 0.50 micro m(-1) x s(-1). These k(cat)/K(m) values are lower but comparable to those of E. coli FKBP22. We propose that a FKBP family protein is involved in cold-adaptation of psychrotrophic bacteria.  相似文献   

7.
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   

8.
The enzyme ornithine carbamoyltransferase (OTCase) of Moritella abyssi (OTCase(Mab)), a new, strictly psychrophilic and piezophilic bacterial species, was purified. OTCase(Mab) displays maximal activity at rather low temperatures (23 to 25 degrees C) compared to other cold-active enzymes and is much less thermoresistant than its homologues from Escherichia coli or thermophilic procaryotes. In vitro the enzyme is in equilibrium between a trimeric state and a dodecameric, more stable state. The melting point and denaturation enthalpy changes for the two forms are considerably lower than the corresponding values for the dodecameric Pyrococcus furiosus OTCase and for a thermolabile trimeric mutant thereof. OTCase(Mab) displays higher K(m) values for ornithine and carbamoyl phosphate than mesophilic and thermophilic OTCases and is only weakly inhibited by the bisubstrate analogue delta-N-phosphonoacetyl-L-ornithine (PALO). OTCase(Mab) differs from other, nonpsychrophilic OTCases by substitutions in the most conserved motifs, which probably contribute to the comparatively high K(m) values and the lower sensitivity to PALO. The K(m) for ornithine, however, is substantially lower at low temperatures. A survey of the catalytic efficiencies (k(cat)/K(m)) of OTCases adapted to different temperatures showed that OTCase(Mab) activity remains suboptimal at low temperature despite the 4.5-fold decrease in the K(m) value for ornithine observed when the temperature is brought from 20 to 5 degrees C. OTCase(Mab) adaptation to cold indicates a trade-off between affinity and catalytic velocity, suggesting that optimization of key metabolic enzymes at low temperatures may be constrained by natural limits.  相似文献   

9.
The activation of human platelets by alpha-thrombin is mediated at least in part by cleavage of protease-activated G-protein-coupled receptors, PAR-1 and PAR-4. Platelet glycoprotein Ibalpha also has a high affinity binding site for alpha-thrombin, and this interaction contributes to platelet activation through a still unknown mechanism. In the present study the hypothesis that GpIbalpha may contribute to platelet activation by modulating the hydrolysis of PAR-1 on the platelet membrane was investigated. Gel-filtered platelets from normal individuals were stimulated by alpha-thrombin, and the kinetics of PAR-1 hydrolysis by enzyme was followed with flow cytometry using an anti-PAR-1 monoclonal antibody (SPAN 12) that recognizes only intact PAR-1 molecules. This strategy allowed measurement of the apparent k(cat)/K(m) value for thrombin hydrolysis of PAR-1 on intact platelets, which was equal to 1.5 +/- 0.1 x 10(7) m(-1) sec(-1). The hydrolysis rate of PAR-1 by thrombin was measured under conditions in which thrombin binding to GpIb was inhibited by different strategies, with the following results. 1) Elimination of GpIbalpha on platelet membranes by mocarhagin treatment reduced the k(cat)/K(m) value by about 6-fold. 2) A monoclonal anti-GpIb antibody reduced the apparent k(cat)/K(m) value by about 5-fold. 3) An oligonucleotide DNA aptamer, HD22, which binds to the thrombin heparin-binding site (HBS) and inhibits thrombin interaction with GpIbalpha, reduced the apparent k(cat)/K(m) value by about 5-fold. 4) Displacement of alpha-thrombin from the binding site on GpIb using PPACK-thrombin reduced the apparent k(cat)/K(m) value by about 5-fold, and 5) mutation at the HBS of thrombin (R98A) caused a 5-fold reduction of the apparent k(cat)/K(m) value of PAR-1 hydrolysis. Altogether these results show that thrombin interaction with GpIb enhances the specificity of thrombin cleavage of PAR-1 on intact platelets, suggesting that GpIb may function as a "cofactor" for PAR-1 activation by thrombin.  相似文献   

10.
The notion of "ground-state destabilization" has been well documented in enzymology. It is the unfavourable interaction (strain) in the enzyme-substrate complex, and increases the k(cat) value without changing the k(cat)/K(m) value. During the course of the investigation on the reaction mechanism of aspartate aminotransferase (AAT), we found another type of strain that is crucial for catalysis: the strain of the distorted internal aldimine in the unliganded enzyme. This strain raises the energy level of the starting state (E+S), thereby reducing the energy gap between E+S and ES(++) and increasing the k(cat)/K(m) value. Further analysis on the reaction intermediates showed that the Michaelis complex of AAT with aspartate contains strain energy due to an unfavourable interaction between the main chain carbonyl oxygen and the Tyr225-aldimine hydrogen-bonding network. This belongs to the classical type of strain. In each case, the strain is reflected in the pK(a) value of the internal aldimine. In the historical explanation of the reaction mechanism of AAT, the shifts in the aldimine pK(a) have been considered to be the driving forces for the proton transfer during catalysis. However, the above findings indicate that the true driving forces are the strain energy inherent to the respective intermediates. We describe here how these strain energies are generated and are used for catalysis, and show that variations in the aldimine pK(a) during catalysis are no more than phenomenological results of adjusting the energy levels of the reaction intermediates for efficient catalysis.  相似文献   

11.
We describe here a new enzyme-coupled assay for the quantitation of d-xylose using readily available enzymes that allows kinetic evaluation of hemicellulolytic enzymes using natural xylooligosaccharide substrates. Hydrogen peroxide is generated as an intermediary analyte, which allows flexibility in the choice of the chromophore or fluorophore used as the final reporter. Thus, we present d-xylose quantitation results for solution-phase assays performed with both the fluorescent reporter resorufin, generated from N-acetyl-3,7-dihydroxyphenoxazine (Amplex Red), and 2,2'-azino-bis(3-ethylbenzothiazoline-6-sulfonate) (ABTS), whose corresponding radical cation has an absorbance maximum at approximately 400 nm. We also describe a useful solid-phase variation of the assay performed with the peroxidase substrate 3,3'-diaminobenzidine tetrahydrochloride, which produces an insoluble brown precipitate. In addition, kinetic parameters for hydrolysis of the natural substrates xylobiose and xylotriose were obtained using this assay for a glycosyl hydrolase family 39 beta-xylosidase from Thermoanaerobacterium sp. strain JW/SL YS485 (Swiss-Prot accession no. O30360). At higher xylobiose substrate concentrations the enzyme showed an increase in the rate indicative of transglycosylation, while for xylotriose marked substrate inhibition was observed. At lower xylobiose concentrations k(cat) was 2.7 +/- 0.4 s(-1), K(m) was 3.3 +/- 0.7 mM, and k(cat)/K(m) was 0.82 +/- 0.21 mM(-1) . s(-1). Nonlinear curve fitting to a substrate inhibition model showed that for xylotriose K(i) was 1.7 +/- 0.1 mM, k(cat) was 2.0 +/- 0.1 s(-1), K(m) was 0.144 +/- 0.011 mM, and k(cat)/K(m) was 14 +/- 1.3 mM(-1) . s(-1).  相似文献   

12.
An alpha-amylase secreted by Pichia burtonii 15-1 isolated from a traditional starter murcha of Nepal, named Pichia burtonii alpha-amylase (PBA), was studied. The gene was cloned and its nucleotide sequence was determined. PBA was deduced to consist of 494 amino acid residues. It shared certain degrees of amino acid sequence identity with other homologous proteins: 60% with Schwanniomyces occidentalis alpha-amylase, 58% with Saccharomycopsis sp. alpha-amylase, and 47% with Taka-amylase A from Aspergillus oryzae. A three-dimensional structural model of PBA generated using the known three-dimensional structure of Taka-amylase A as a template suggested high structural similarity between them. Kinetic analysis revealed that the K(m) values of PBA were lower than those of Taka-amylase A for the oligosaccharides. Although the k(cat) values of PBA were lower than those of Taka-amylase A for the oligosaccharide substrates, the k(cat)/K(m) values of PBA were higher.  相似文献   

13.
In the blood coagulation cascade, thrombin cleaves fibrinopeptides A and B from fibrinogen revealing sites for fibrin polymerization that lead to insoluble clot formation. Factor XIII stabilizes this clot by catalyzing the formation of intermolecular cross-links in the fibrin network. Thrombin activates the Factor XIII a(2) dimer by cleaving the Factor XIII activation peptide segment at the Arg(37)-Gly(38) peptide bond. Using a high performance liquid chromatography assay, the kinetic constants K(m), k(cat), and k(cat)/K(m) were determined for thrombin hydrolysis of fibrinogen Aalpha-(7-20), Factor XIII activation peptide-(28-41), and Factor XIII activation peptide-(28-41) with a Val(34) to Leu substitution. This Val to Leu mutation has been correlated with protection from myocardial infarction. In the absence of fibrin, the Factor XIII activation peptide-(28-41) exhibits a 10-fold lower k(cat)/K(m) value than fibrinogen Aalpha-(7-20). With the Factor XIII V34L mutation, decreases in K(m) and increases in k(cat) produce a 6-fold increase in k(cat)/K(m) relative to the wild-type Factor XIII sequence. A review of the x-ray crystal structures of known substrates and inhibitors of thrombin leads to a hypothesis that the new Leu generates a peptide with more extensive interactions with the surface of thrombin. As a result, the Factor XIII V34L is proposed to be susceptible to wasteful conversion of zymogen to activated enzyme. Premature depletion may provide cardioprotective effects.  相似文献   

14.
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   

15.
Vocadlo DJ  Wicki J  Rupitz K  Withers SG 《Biochemistry》2002,41(31):9727-9735
The catalytic mechanism of Thermoanaerobacterium saccharolyticum beta-xylosidase (XynB) from family 39 of glycoside hydrolases has been subjected to a detailed kinetic investigation using a range of substrates. The enzyme exhibits a bell-shaped pH dependence of k(cat)/K(m), reflecting apparent pK(a) values of 4.1 and 6.8. The k(cat) and k(cat)/K(m) values for a series of aryl xylosides have been measured and used to construct two Br?nsted plots. The plot of log(k(cat)/K(m)) against the pK(a) of the leaving group reveals a significant correlation (beta(lg) = -0.97, r(2) = 0.94, n = 8), indicating that fission of the glycosidic bond is significantly advanced in the transition state leading to the formation of the xylosyl-enzyme intermediate. The large negative value of the slope indicates that there is relatively little proton donation to the glycosidic oxygen in the transition state. A biphasic, concave-downward plot of log(k(cat)) against pK(a) provides good evidence for a two-step double-displacement mechanism involving a glycosyl-enzyme intermediate. For activated leaving groups (pK(a) < 9), the breakdown of the xylosyl-enzyme intermediate is the rate-determining step, as indicated by the absence of any effect of the pK(a) of the leaving group on log(k(cat)) (beta(lg) approximately 0). However, a strong dependence of the first-order rate constant on the pK(a) value of relatively poor leaving groups (pK(a) > 9) suggests that the xylosylation step is rate-determining for these substrates. Support for the dexylosylation chemical step being rate-determining for activated substrates comes from nucleophilic competition experiments in which addition of dithiothreitol results in an increase in turnover rates. Normal secondary alpha-deuterium kinetic isotope effects ((alpha-D)(V) or (alpha-D)(V/K) = 1.08-1.10) for three different substrates of widely varying pK(a) value (5.15-9.95) have been measured and these reveal that the transition states leading to the formation and breakdown of the intermediate are similar and both steps involve rehybridization of C1 from sp(3) to sp(2). These results are consistent only with "exploded" transition states, in which the saccharide moiety bears considerable positive charge, and the intermediate is a covalent acylal-ester where C1 is sp(3) hybridized.  相似文献   

16.
To obtain a clearer understanding of the forces involved in transition state stabilization by Escherichia coli cytidine deaminase, we investigated the thermodynamic changes that accompany substrate binding in the ground state and transition state for substrate hydrolysis. Viscosity studies indicate that the action of cytidine deaminase is not diffusion-limited. Thus, K(m) appears to be a true dissociation constant, and k(cat) describes the chemical reaction of the ES complex, not product release. Enzyme-substrate association is accompanied by a loss of entropy and a somewhat greater release of enthalpy. As the ES complex proceeds to the transition state (ES), there is little further change in entropy, but heat is taken up that almost matches the heat that was released with ES formation. As a result, k(cat)/K(m) (describing the overall conversion of the free substrate to ES is almost invariant with changing temperature. The free energy barrier for the enzyme-catalyzed reaction (k(cat)/K(m)) is much lower than that for the spontaneous reaction (k(non)) (DeltaDeltaG = -21.8 kcal/mol at 25 degrees C). This difference, which also describes the virtual binding affinity of the enzyme for the activated substrate in the transition state (S), is almost entirely enthalpic in origin (DeltaDeltaH = -20.2 kcal/mol), compatible with the formation of hydrogen bonds that stabilize the ES complex. Thus, the transition state affinity of cytidine deaminase increases rapidly with decreasing temperature. When a hydrogen bond between Glu-91 and the 3'-hydroxyl moiety of cytidine is disrupted by truncation of either group, k(cat)/K(m) and transition state affinity are each reduced by a factor of 10(4). This effect of mutation is entirely enthalpic in origin (DeltaDeltaH approximately 7.9 kcal/mol), somewhat offset by a favorable change in the entropy of transition state binding. This increase in entropy is attributed to a loss of constraints on the relative motions of the activated substrate within the ES complex. In an Appendix, some objections to the conventional scheme for transition state binding are discussed.  相似文献   

17.
The phosphoglucomutase gene from a wild type Fusarium oxysporum strain (F3), was homologously expressed, under the control of the constitutive promoter of gpdA of Aspergillus nidulans. The transformant produced elevated levels of phosphoglucomutase activity compared to the wild type, a fact that facilitated the subsequent purification procedure. The enzyme (FoPGM) was purified to homogeneity applying three anion exchange and one gel filtration chromatography steps. The native enzyme revealed a monomeric structure with a molecular mass of 60 kDa, while the isoelectric point was 3.5. FoPGM was active in pH ranged from 6.0 to 8.0, with an optimum using 3-(N-morpholino)propanesulfonic acid buffer at 7.0, while loss of activity was observed when phosphate buffer was used in the above mentioned pH range. The optimal temperature for activity was 45°C but the enzyme became unstable at temperatures above 40°C. FoPGM requires the presence of a divalent cation for its function with maximum activity being obtained with Co(2+). The apparent K(m) for Co(2+) was found to be 10 μM. The enzyme was also active with other divalent metal ions such as Mn(2+), Mg(2+), Ni(2+) and Ca(2+) but to a lesser extent. The following kinetic constants were determined: v(max), 0.74 μmol mg(protein)(-1)min(-1); k(cat), 44.2 min(-1); K(m)(G1P), 0.10mM; K(m)(G1,6 diP), 1.03 μM; k(cat)/K(m)(G1P), 443 mM(-1)min(-1) and k(cat)/K(m)(G1,6 diP), 42,860 mM(-1)min(-1). The enzyme was considered to follow a Ping Pong substituted enzyme or enzyme isomerization mechanism.  相似文献   

18.
To elucidate the functions of the surface loops of VIIa, we prepared two mutants, VII-30 and VII-39. The VII-30 mutant had all of the residues in the 99 loop replaced with those of trypsin. In the VII-39 mutant, both the 99 and 170 loops were replaced with those of trypsin. The k(cat)/K(m) value for hydrolysis of the chromogenic peptidyl substrate S-2288 by VIIa-30 (103 mm(-)1s(-)1) was 3-fold higher than that of wild-type VIIa (30.3 mm(-)1 s(-)1) in the presence of soluble tissue factor (sTF). This enhancement was due to a decrease in the K(m) value but not to an increase in the k(cat) value. On the other hand, the k(cat)/K(m) value for S-2288 hydrolysis by VIIa-39 (17.9 mm(-)1 s(-)1) was 18-fold higher than that of wild-type (1.0 mm(-)1 s(-)1) in the absence of sTF, and the value was almost the same as that of wild-type measured in the presence of sTF. This enhancement was due to not only a decrease in the K(m) value but also to an increase in the k(cat) value. These results were in good agreement with their susceptibilities to a subsite 1-directed serine protease inhibitor. In our previous paper (Soejima, K., Mizuguchi, J., Yuguchi, M., Nakagaki, T., Higashi, S., and Iwanaga, S. (2001) J. Biol. Chem. 276, 17229-17235), the replacement of the 170 loop of VIIa with that of trypsin induced a 10-fold enhancement of the k(cat) value for S-2288 hydrolysis as compared with that of wild-type VIIa in the absence of sTF. These results suggested that the 99 and the 170 loop structures of VIIa independently affect the K(m) and k(cat) values, respectively. Furthermore, we studied the effect of mutations on proteolytic activity toward S-alkylated lysozyme as a macromolecular substrate and the activation of natural macromolecular substrate factor X.  相似文献   

19.
A series of Glu(pNA)-containing peptides was designed to determine the activity of the transglutaminase factor XIIIa at 405 nm due to p-nitroaniline release. The most suitable substrate properties were found for peptides containing the Glu(pNA) residue in the second position from the N terminus. For the best substrate 12 (H-Tyr-Glu(pNA)-Val-Lys-Val-Ile-Gly-NH(2)), a k(cat)/K(m) value of 3531 s(-1)M(-1) was found. Although the k(cat)/K(m) values of the Glu(pNA) peptides are more than 100-fold reduced compared with the previously reported cleavage of natural glutamine-containing substrates such as α(2)-antiplasmin and β-casein, these chromogenic substrates can be useful tools for convenient determination of FXIII-A(2)* activity e.g., for in vitro inhibitor screening. As an example, peptide 12 was used to characterize the inhibition of FXIII-A(2)* by the well-known irreversible inhibitor iodoacetic acid.  相似文献   

20.
A new thermostable dipeptidase gene was cloned from the thermophile Brevibacillus borstelensis BCS-1 by genetic complementation of the D-Glu auxotroph Escherichia coli WM335 on a plate containing D-Ala-D-Glu. Nucleotide sequence analysis revealed that the gene included an open reading frame coding for a 307-amino-acid sequence with an M(r) of 35,000. The deduced amino acid sequence of the dipeptidase exhibited 52% similarity with the dipeptidase from Listeria monocytogenes. The enzyme was purified to homogeneity from recombinant E. coli WM335 harboring the dipeptidase gene from B. borstelensis BCS-1. Investigation of the enantioselectivity (E) to the P(1) and P(1)' site of Ala-Ala revealed that the ratio of the specificity constant (k(cat)/K(m)) for L-enantioselectivity to the P(1) site of Ala-Ala was 23.4 +/- 2.2 [E = (k(cat)/K(m))(L,D)/(k(cat)/K(m))(D,D)], while the D-enantioselectivity to the P(1)' site of Ala-Ala was 16.4 +/- 0.5 [E = (k(cat)/K(m))(L,D)/(k(cat)/K(m))(L,L)] at 55 degrees C. The enzyme was stable up to 55 degrees C, and the optimal pH and temperature were 8.5 and 65 degrees C, respectively. The enzyme was able to hydrolyze L-Asp-D-Ala, L-Asp-D-AlaOMe, Z-D-Ala-D-AlaOBzl, and Z-L-Asp-D-AlaOBzl, yet it could not hydrolyze D-Ala-L-Asp, D-Ala-L-Ala, D-AlaNH(2), and L-AlaNH(2.) The enzyme also exhibited beta-lactamase activity similar to that of a human renal dipeptidase. The dipeptidase successfully synthesized the precursor of the dipeptide sweetener Z-L-Asp-D-AlaOBzl.  相似文献   

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