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 298 毫秒
1.
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   

2.
Protease-substrate interactions are governed by a variety of structural features. Although the substrate sequence specificities of numerous proteases have been established, "topological specificities," whereby proteases may be classified based on recognition of distinct three-dimensional structural motifs, have not. The aggrecanase members of the ADAMTS (a disintegrin and metalloproteinase with thrombospondin motifs) family cleave a variety of proteins but do not seem to possess distinct sequence specificities. In the present study, the topological substrate specificity of ADAMTS-4 (aggrecanase-1) was examined using triple-helical or single-stranded poly(Pro) II helical peptides. Substrate topology modulated the affinity and sequence specificity of ADAMTS-4 with K(m) values indicating a preference for triple-helical structure. In turn, non-catalytic ADAMTS-4 domains were critical for hydrolysis of triple-helical and poly(Pro) II helical substrates. Comparison of ADAMTS-4 with MMP-1 (collagenase 1), MMP-13 (collagenase 3), trypsin, and thermolysin using triple-helical peptide (THP) and single-stranded peptide (SSP) substrates demonstrated that all five proteases possessed efficient "triple-helical peptidase" activity and fell into one of two categories: (k(cat)/K(m))(SSP) > (k(cat)/K(m))(THP) (thermolysin, trypsin, and MMP-13) or (k(cat)/K(m))(THP) > or = (k(cat)/K(m))(SSP) and (K(m))(SSP) > (K(m))(THP) (MMP-1 and ADAMTS-4). Overall these results suggest that topological specificity may be a guiding principle for protease behavior and can be utilized to design specific substrates and inhibitors. The triple-helical and single-stranded poly(Pro) II helical peptides represent the first synthetic substrates successfully designed for aggrecanases.  相似文献   

3.
The effects of calcium ions on hydrolysis of low molecular weight substrates catalyzed by different forms of enteropeptidase were studied. A method for determining activity of truncated enteropeptidase preparations lacking a secondary trypsinogen binding site and displaying low activity towards trypsinogen was developed using N-alpha-benzyloxycarbonyl-L-lysine thiobenzyl ester (Z-Lys-S-Bzl). The kinetic constants for hydrolysis of this substrate at pH 8.0 and 25 degrees C were determined for natural enteropeptidase (K(m) 59.6 microM, k(cat) 6660 min(-1), k(cat)/K(m) 111 microM(-1) x min(-1)), as well as for enteropeptidase preparation with deleted 118-783 fragment of the heavy chain (K(m) 176.9 microM, k(cat) 6694 min(-1), k(cat)/K(m) 37.84 microM(-1) x min(-1)) and trypsin (K(m) 56.0 microM, k(cat) 8280 min(-1), k(cat)/K(m) 147.86 microM(-1) x min(-1)). It was shown that the enzymes with trypsin-like primary active site display similar hydrolysis efficiency towards Z-Lys-S-Bzl. Calcium ions cause 3-fold activation of hydrolysis of the substrates of general type GD(4)K-X by the natural full-length enteropeptidase. In contrast, the hydrolysis of substrates with one or two Asp/Glu residues at P2-P3 positions is slightly inhibited by Ca2+. In the case of enteropeptidase light chain as well as the enzyme containing the truncated heavy chain (466-800 fragment), the activating effect of calcium ions was not detected for all the studied substrates. The results of hydrolysis experiments with synthetic enteropeptidase substrates GD(4)K-F(NO(2))G, G(5)DK-F(NO(2))G (where F(NO(2)) is p-nitrophenyl-L-phenylalanine residue), and GD(4)K-Nfa (where Nfa is beta-naphthylamide) demonstrate the possibility of regulation of undesired side hydrolysis using natural full-length enteropeptidase for processing chimeric proteins by means of calcium ions.  相似文献   

4.
Amino and carboxyl termini of the bifunctional enzyme Fru 6-P, 2-kinase:Fru 2,6-bisphosphatase regulate the relative activities of the kinase/phosphatase. The N-terminus of the rat liver bifunctional enzyme is highly basic, containing a protein kinase A phosphorylation site that regulates these enzyme activities in a reciprocal manner. To determine the role of charged residues in the N-terminal peptide, mutant enzymes were constructed in which these residues were altered to residues carrying opposite charges, and the effect on the catalytic properties, thermal lability, and susceptibility to trypsin digestion and phosphorylation by protein kinase A was determined. Most of these mutations decreased k(cat)/K(ATP) and/or k(cat)/K(Fru) (6-P) of the kinase and increased k(cat)/K(Fru 2,6-P2) of the phosphatase. These mutant enzymes were more susceptible to trypsin digestion, phosphorylation by protein kinase A, and thermal inactivation. In general, the effect was greater with amino acid residues located more distant from the N-terminus. The resulting changes were not as large as observed with the phosphorylated enzyme. Mutation of Ser22 to Pro produced large changes in the kinetic properties comparable to those of phosphorylation, suggesting that the flexible region of the N-terminus containing five serines (Ser20 to S24) is essential for the enzyme activities. These results indicated that the charged residues as well as Ser20-Ser24 in the N-terminus of the liver Fru 6-P,2-kinase:Fru 2,6-Pase are essential in the allosteric regulation and probably involved in interactions with the catalytic domains that induce a conformation that has high Fru 6-P,2-kinase and low Fru 2,6-Pase activities. Any disruption of this N-terminal interaction results in inhibition of the kinase and activation of the phosphatase.  相似文献   

5.
The Staphylococcus aureus transpeptidase Sortase A (SrtA) anchors virulence and colonization-associated surface proteins to the cell wall. SrtA selectively recognizes a C-terminal LPXTG motif, whereas the related transpeptidase Sortase B (SrtB) recognizes a C-terminal NPQTN motif. In both enzymes, cleavage occurs after the conserved threonine, followed by amide bond formation between threonine and the pentaglycine cross-bridge of cell wall peptidoglycan. Genetic and biochemical studies strongly suggest that SrtA and SrtB exhibit exquisite specificity for their recognition motifs. To better understand the origins of substrate specificity within these two isoforms, we used sequence and structural analysis to predict residues and domains likely to be involved in conferring substrate specificity. Mutational analyses and domain swapping experiments were conducted to test their function in substrate recognition and specificity. Marked changes in the specificity profile of SrtA were obtained by replacing the beta6/beta7 loop in SrtA with the corresponding domain from SrtB. The chimeric beta6/beta7 loop swap enzyme (SrtLS) conferred the ability to acylate NPQTN-containing substrates, with a k(cat)/K(m)(app) of 0.0062 +/- 0.003 m(-1) s(-1). This enzyme was unable to perform the transpeptidation stage of the reaction, suggesting that additional domains are required for transpeptidation to occur. The overall catalytic specificity profile (k(cat)/K(m)(app)(NPQTN)/k(cat)/K(m)(app)(LPETG)) of SrtLS was altered 700,000-fold from SrtA. These results indicate that the beta6/beta7 loop is an important site for substrate recognition in sortases.  相似文献   

6.
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   

7.
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   

8.
Conjugate DNAzymes with polyamines and peptides were successfully prepared by solid phase fragment condensation (SPFC) and showed up to 4.2 times higher catalytic efficiency (k(cat)/K(m)) and enhanced tolerance against DNase 1digestion. To be pointed out, intracellular localization of DNAzymes could be controlled by conjugated with naturally occurring signal peptides responsible for nuclear cytoplasmic transport of proteins.  相似文献   

9.
Purification and characterization of active recombinant rat kallikrein rK9.   总被引:3,自引:0,他引:3  
The rat tissue kallikrein rK9 is most abundant in the submandibular gland and the prostate. It has been successfully expressed in the Pichia pastoris yeast expression system. A full-length cDNA coding for the mature rK9 was fused in frame with yeast alpha-factor cDNA. The fusion protein was secreted into the medium with high yield without being processed by the yeast KEX2 signal peptidase. Mature rK9 was efficiently released from the fusion protein by trypsin and was purified to homogeneity by one-step affinity chromatography using soya bean trypsin inhibitor (SBTI) as affinity ligand. The identity of the recombinant enzyme was checked by N-terminal amino acid sequencing, Western blot analysis and kinetic studies. The dual trypsin- and chymotrypsin-like enzymatic specificity of rK9 was assessed by determining specificity constants (k(cat)/K(m)) for the hydrolysis of fluorogenic substrates, the peptide sequences of which were derived from proparathyroid hormone (pro-PTH) and from semenogelin-I. Our results confirmed the presence of an extended binding site in the rK9 active site. We also identified a far more sensitive substrate of this enzyme than those previously described, Abz-VKKRSARQ-EDDnp, which was hydrolysed with a catalytic efficiency k(cat)/K(m) of 420000 M(-1)s(-1). Finally, we showed that four of the five major proteins contained in secretions of rat seminal vesicles were rapidly degraded by recombinant rK9.  相似文献   

10.
Osteoclasts and macrophages express high amounts of tartrate-resistant acid phosphatase (TRACP), an enzyme with unknown biological function. TRACP contains a disulfide bond, a protease-sensitive loop peptide, and a redox-active iron that can catalyze formation of reactive oxygen species (ROS). We studied the effects of proteolytic cleavage by trypsin, reduction of the disulfide bond by beta-mercaptoethanol, and reduction of the redox-active iron by ascorbate on the phosphatase and ROS-generating activity of baculovirus-generated recombinant human TRACP. Ascorbate alone and trypsin in combination with beta-mercaptoethanol increased k(cat)/K(m) of the phosphatase activity seven- to ninefold. The pH-optimum was changed from 5.4-5.6 to 6.2-6.4 by ascorbate and trypsin cleavage. Trypsin cleavage increased k(cat)/K(m) of the ROS-generating activity 2.5-fold without affecting the pH-optimum (7.0). These results suggest that the protease-sensitive loop peptide, redox-active iron, and disulfide bond are important regulatory sites in TRACP, and that the phosphatase and ROS-generating activity are performed with different reaction mechanisms.  相似文献   

11.
Based on recent directed evolution of P450 2B1, six P450 2B11 mutants at three positions were created in an N-terminal modified construct termed P450 2B11dH and characterized for enzyme catalysis using five substrates. Mutant I209A demonstrated a 3.2-fold enhanced k(cat)/K(m) for 7-ethoxy-4-trifluoromethylcourmarin O-deethylation, largely due to a dramatic decrease in K(m) (0.72 microM vs. 18 microM). I209A also demonstrated enhanced selectivity for testosterone 16beta-hydroxylation over 16alpha-hydroxylation. In contrast, V183L showed a 4-fold increased k(cat) for 7-benzyloxyresorufin debenzylation and a 4.7-fold increased k(cat)/K(m) for testosterone 16alpha-hydroxylation. V183L also displayed a 1.7-fold higher k(cat)/K(m) than P450 2B11dH with the anti-cancer prodrugs cyclophosphamide and ifosfamide, resulting from a approximately 4-fold decrease in K(m). Introduction of the V183L mutation into full-length P450 2B11 did not enhance the k(cat)/K(m). Overall, the re-engineered P450 2B11dH enzymes exhibited enhanced catalytic efficiency with several substrates including the anti-cancer prodrugs.  相似文献   

12.
Tryptic digestion followed by identification using mass spectrometry is an important step in many proteomic studies. Here, we describe the preparation of immobilized, acetylated trypsin for enhanced digestion efficacy in integrated protein analysis platforms. Complete digestion of cytochrome c was obtained with two types of modified-trypsin beads with a contact time of only 4 s, while corresponding unmodified-trypsin beads gave only incomplete digestion. The digestion rate of myoglobin, a protein known to be rather resistant to proteolysis, was not altered by acetylating trypsin and required a buffer containing 35% acetonitrile to obtain complete digestion. The use of acetylated-trypsin beads led to fewer interfering tryptic autolysis products, indicating an increased stability of this modified enzyme. Importantly, the modification did not affect trypsin's substrate specificity, as the peptide map of myoglobin was not altered upon acetylation of immobilized trypsin. Kinetic digestion experiments in solution with low-molecular-weight substrates and cytochrome c confirmed the increased catalytic efficiency (lower K(M) and higher k(cat)) and increased resistance to autolysis of trypsin upon acetylation. Enhancement of catalytic efficiency was correlated with the number of acetylations per molecule. The favorable properties of the new chemically modified trypsin reactor should make it a valuable tool in automated protein analysis systems.  相似文献   

13.
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   

14.
In contrast to the prevalent view in the literature hitherto, the present study shows that pancreatic trypsin can activate human promatrix metalloproteinase-2 (proMMP-2). It is shown that trypsin's ability to activate proMMP-2 is dependent on various environmental factors such as the level of exogenously added Ca(2+) and Brij-35, temperature, as well as trypsin concentration. The activation occurred as a sequential processing of the proenzyme, initially generating an active 62kDa species. This was followed by successive truncation of the C-terminal domain, giving rise to active species of 56kDa, 52kDa and 50kDa. Tissue inhibitor of matrix metalloproteinases-2 (TIMP-2) prevented the trypsin-mediated C-terminal truncation, without affecting the generation of the 62kDa species, while the presence of EDTA increased the rate of the trypsin-mediated activation of proMMP-2. MALDI-TOF MS analysis of the 50kDa form indicated that trypsin generated active forms with either Lys87 or Trp90 as the N-terminal residue and Arg538 as a C-terminal residue. The trypsin-activated MMP-2 was active in solution against both synthetic and physiologic substrates, and the steady-state kinetic coefficients k(cat), K(m) and k(cat)/K(m) were determined for the enzyme activated either by APMA, membrane-type 1 matrix metalloproteinase (MT1-MMP) or trypsin. The trypsin-activated MMP-2 exhibited slightly lower k(cat) and k(cat)/K(m) values as well as a slightly higher K(i) value against TIMP-1 compared to the enzyme activated by APMA or MT1-MMP. Docking studies of TIMP-1 revealed that the slightly weaker binding of the inhibitor to the trypsin-activated MMP-2 could be attributed to its shorter N terminus (Lys87/Trp90 versus Tyr81), as Phe83 and Arg86 interacted directly with the inhibitor. Our results suggest that the trypsin-activated MMP-2 possesses the catalytic and regulatory potential to be of significance in vivo.  相似文献   

15.
A series of 16 bovine pancreatic trypsin inhibitor variants mutated at the P(1) position of the binding loop and seven tetrapeptide p-nitroanilide (pNa) substrates of the general formula: suc-Ala-Ala-Pro-Aaa-pNa (where Aaa denotes either: Phe, Arg, Lys, Leu, Met, Nva, Nle) were used to investigate the influence of high salt concentration on the activity of bovine chymotrypsin. The increase of the association constant (K(a)) and the specificity index (k(cat)/K(m)) in the presence of 3 M NaCl highly depends on the chemical nature of the residue at the P(1) position. The highest increase was observed for inhibitors/substrates containing the basic side chains at this site. Surprisingly, for the remaining 13 residues the observed salt effect is not correlated with any side chain properties. In particular, there is a lack of correlation between the accessible non-polar surface area and the magnitude of the salt effect. It suggests that salt-induced increase of the K(a) and k(cat)/K(m) values is not caused by the enhancement of the hydrophobic interactions in chymotrypsin-inhibitor/substrate complex. Moreover, the increase of the K(a) and k(cat)/K(m) values occurs only in the presence of Na(+) ions, while K(+) and Li(+) ions do not change the activity of chymotrypsin. Additionally, the activities of two other proteinases: bovine trypsin and Streptomyces griseus proteinase B were tested in the presence of 3 M NaCl using their specific substrates. The activity of both enzymes was almost not affected by the presence of high NaCl concentration.  相似文献   

16.
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   

17.
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   

18.
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   

19.
Human fibroblast activation protein (FAP), an integral membrane serine protease, was produced in insect cells as a hexa-His-tagged protein using a recombinant baculovirus expression system. Two isoforms of FAP, glycosylated and nonglycosylated, were identified by Western blotting using an anti-His-tag antibody and separated by lectin chromatography. The glycosylated FAP was purified to near homogeneity using immobilized metal affinity chromatography and was shown to have both postprolyl dipeptidyl peptidase and postgelatinase activities. In contrast, the nonglycosylated isoform demonstrated no detectable gelatinase activity by either zymography or a fluorescence-based gelatinase activity assay. The kinetic parameters of the dipeptidyl peptidase activity for glycosylated FAP were determined using dipeptide Ala-Pro-7-amino-trifluoromethyl-coumarin as the substrate. The k(cat) is 2.0 s(-1) and k(cat)/K(m) is 1.0 x 10(4) M(-1) s(-1) at pH 8.5. The pH dependence of k(cat) reveals two ionization groups with pK(a1) of 7.0 and pK(a2) of 11.0. The pH profile of k(cat)/K(m) yields similar results with pK(a1) 6.2 and pK(a2) 11.0. The neutral pK(a1) is associated with His at the active site. The basic pK(a2) might be contributed from an ionization group that is not involved directly in catalysis, instead associated with the stability of the active site structure.  相似文献   

20.
Ribitol dehydrogenase from Zymomonas mobilis (ZmRDH) catalyzes the conversion of ribitol to d-ribulose and concomitantly reduces NAD(P)(+) to NAD(P)H. A systematic approach involving an initial sequence alignment-based residue screening, followed by a homology model-based screening and site-directed mutagenesis of the screened residues, was used to study the molecular determinants of the cofactor specificity of ZmRDH. A homologous conserved amino acid, Ser156, in the substrate-binding pocket of the wild-type ZmRDH was identified as an important residue affecting the cofactor specificity of ZmRDH. Further insights into the function of the Ser156 residue were obtained by substituting it with other hydrophobic nonpolar or polar amino acids. Substituting Ser156 with the negatively charged amino acids (Asp and Glu) altered the cofactor specificity of ZmRDH toward NAD(+) (S156D, [k(cat)/K(m)(,NAD)]/[k(cat)/K(m)(,NADP)] = 10.9, where K(m)(,NAD) is the K(m) for NAD(+) and K(m)(,NADP) is the K(m) for NADP(+)). In contrast, the mutants containing positively charged amino acids (His, Lys, or Arg) at position 156 showed a higher efficiency with NADP(+) as the cofactor (S156H, [k(cat)/K(m)(,NAD)]/[k(cat)/K(m)(,NADP)] = 0.11). These data, in addition to those of molecular dynamics and isothermal titration calorimetry studies, suggest that the cofactor specificity of ZmRDH can be modulated by manipulating the amino acid residue at position 156.  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