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 62 毫秒
1.
Enteropeptidase is a heterodimeric type II membrane protein of the brush border of duodenal enterocytes. In this location, enteropeptidase cleaves and activates trypsinogen, thereby initiating the activation of other intestinal digestive enzymes. Recombinant bovine enteropeptidase was sorted directly to the apical surface of polarized Madin-Darby canine kidney cells. Replacement of the cytoplasmic and signal anchor domains with a cleavable signal peptide (mutant proenteropeptidase lacking the amino-terminal signal anchor domain (dSA-BEK)) caused apical secretion. The additional amino-terminal deletion of a mucin-like domain (HL-BEK) resulted in secretion both apically and basolaterally. Further deletion of the noncatalytic heavy chain (L-BEK) resulted in apical secretion. Thus enteropeptidase appears to have at least three distinct sorting signals as follows: the light chain (L-BEK) directs apical sorting, addition of most of the heavy chain (HL-BEK) inhibits apical sorting, and addition of the mucin-like domain (dSA-BEK) restores apical sorting. Inhibition of N-linked glycosylation with tunicamycin or disruption of microtubules with colchicine caused L-BEK to be secreted equally into apical and basolateral compartments, whereas brefeldin A caused basolateral secretion of L-BEK. Full-length BEK was not found in detergent-resistant raft domains of Madin-Darby canine kidney cells or baby hamster kidney cells. These results suggest apical sorting of enteropeptidase depends on N-linked glycosylation of the serine protease domain and an amino-terminal segment that includes an O-glycosylated mucin-like domain and three potential N-glycosylation sites. In contrast to many apically targeted proteins, enteropeptidase does not form detergent-resistant associations with sphingolipid-cholesterol rafts.  相似文献   

2.
Pseudomonas carboxyl proteinase (PCP), isolated from Pseudomonas sp. 101, and Xanthomonas carboxyl proteinase (XCP), isolated from Xanthomonas sp. T-22, are the first and second examples of unique carboxyl proteinases [EC 3.4.23.33] which are insensitive to aspartic proteinase inhibitors, such as pepstatin, diazoacetyl-DL-norleucine methylester, and 1,2-epoxy-3(p-nitrophenoxy)propane. The substrate specificities of PCP and XCP were studied using a series of synthetic chromogenic peptide substrates with the general structure, P5-P4-P3-P2-Phe-Nph-P2'-P3' (P5, P4, P3, P2, P2', P3': a variety of amino acids, Nph is p-nitro-L-phenylalanine, and the Phe-Nph bond is cleaved). PCP and XCP were shown to hydrolyze a synthetic substrate, Lys-Pro-Ala-Leu-Phe-Nph-Arg-Leu, most effectively among 28 substrates. The kinetic parameters of this peptide for PCP were Km = 6.3 microM, Kcat = 51.4 s-1, and kcat/Km = 8.16 microM-1.s-1. The kinetic parameters for XCP were Km = 3.6 microM, kcat = 52.2 s-1, and kcat/Km = 14.5 microM-1.s-1. PCP showed a stricter substrate specificity than XCP. That is, the specificity constant (kcat/Km) of each substrate for PCP was in general < 0.5 microM-1.s-1, but was drastically improved by the replacement of Lys by Leu at the P2 position. On the other hand, XCP showed a less stringent substrate specificity, with most of the peptides exhibiting reasonable kcat/Km values (> 1.0 microM-1.s-1). Thus it was found that the substrate specificities of PCP and XCP differ considerably, in spite of the high similarity in their primary structures. In addition, tyrostatin was found to be a competitive inhibitor for XCP, with a Ki value of 2.1 nM, as well as for PCP (Ki = 2.6 nM).  相似文献   

3.
Ricin A-chain (RTA) catalyzes the depurination of a single adenine at position 4324 of 28S rRNA in a N-ribohydrolase reaction. The mechanism and specificity for RTA are examined using RNA stem-loop structures of 10-18 nucleotides which contain the required substrate motif, a GAGA tetraloop. At the optimal pH near 4.0, the preferred substrate is a 14-base stem-loop RNA which is hydrolyzed at 219 min-1 with a kcat/Km of 4.5 x 10(5) M-1 s-1 under conditions of steady-state catalysis. Smaller or larger stem-loop RNAs have lower kcat values, but all have Km values of approximately 5 microM. Both the 10- and 18-base substrates have kcat/Km near 10(4) M-1 s-1. Covalent cross-linking of the stem has a small effect on the kinetic parameters. Stem-loop DNA (10 bases) of the same sequence is also a substrate with a kcat/Km of 0.1 that for RNA. Chemical mechanisms for enzymatic RNA depurination reactions include leaving group activation, stabilization of a ribooxocarbenium transition state, a covalent enzyme-ribosyl intermediate, and ionization of the 2'-hydroxyl. A stem-loop RNA with p-nitrophenyl O-riboside at the depurination site is not a substrate, but binds tightly to the enzyme (Ki = 0.34 microM), consistent with a catalytic mechanism of leaving group activation. The substrate activity of stem-loop DNA eliminates ionization of the 2'-hydroxyl as a mechanism. Incorporation of the C-riboside formycin A at the depurination site provides an increased pKa of the adenine analogue at N7. Binding of this analogue (Ki = 9.4 microM) is weaker than substrate which indicates that the altered pKa at this position is not an important feature of transition state recognition. Stem-loop RNA with phenyliminoribitol at the depurination site increases the affinity substantially (Ki = 0.18 microM). The results are consistent with catalysis occurring by leaving group protonation at ring position(s) other than N7 leading to a ribooxocarbenium ion transition state. Small stem-loop RNAs have been identified with substrate activity within an order of magnitude of that reported for intact ribosomes.  相似文献   

4.
HTLV-I is an oncogenic retrovirus that is associated with adult T-cell leukemia. HTLV-I protease and HTLV-I protease fused to a deca-histidine containing leader peptide (His-protease) have been cloned, expressed, and purified. The refolded proteases were active and exhibited nearly identical enzymatic activities. To begin to characterize the specificity of HTLV-I, we measured protease cleavage of peptide substrates and inhibition by protease inhibitors. HTLV-I protease cleavage of a peptide representing the HTLV-I retroviral processing site P19/24 (APQVLPVMHPHG) yielded Km and kcat values of 470 microM and 0.184 s-1 while cleavage of a peptide representing the processing site P24/15 (KTKVLVVQPK) yielded Km and kcat values of 310 microM and 0.0060 s-1. When the P1' proline of P19/24 was replaced with p-nitro-phenylalanine (Nph), the ability of HTLV-I protease to cleave the substrate (APQVLNphVMHPL) was improved. Inhibition of HTLV-I protease and His-protease by a series of protease inhibitors was also tested. It was found that the Ki values for inhibition of HTLV-I protease and His-protease by a series of pepsin inhibitors ranged from 7 nM to 10 microM, while the Ki values of a series of HIV-1 protease inhibitors ranged from 6 nM to 127 microM. In comparison, the Ki values for inhibition of pepsin by the pepsin inhibitors ranged from 0.72 to 19.2 nM, and the Ki values for inhibition of HIV-1 protease by the HIV protease inhibitors ranged from 0.24 nM to 1.0 microM. The data suggested that the substrate binding site of HTLV-I protease is different from the substrate binding sites of pepsin and HIV-1 protease, and that currently employed HIV-1 protease inhibitors would not be effective for the treatment of HTLV-I infections.  相似文献   

5.
The transmembrane PTPase HPTP beta differs from its related family members in having a single rather than a tandemly duplicated cytosolic catalytic domain. We have expressed the 354-amino acid, 41-kDa human PTP beta catalytic fragment in Escherichia coli, purified it, and assessed catalytic specificity with a series of pY peptides. HPTP beta shows distinctions from the related LAR PTPase and T cell CD45 PTPase domains: it recognizes phosphotyrosyl peptides of 9-11 residues from lck, src, and PLC gamma with Km values of 2, 4, and 1 microM, some 40-200-fold lower than the other two PTPases. With kcat values of 30-205 s-1, the catalytic efficiency, kcat/Km, of the HPTP beta 41-kDa catalytic domain is very high, up to 5.7 x 10(7) M-1 s-1. The peptides corresponding to PLC gamma (766-776) and EGFR (1,167-1,177) phosphorylation sites were used for structural variation to assess pY sequence context recognition by HPTP beta catalytic domain. While exchange of the alanine residue at the +2 position of the PLC gamma (Km of 1 microM) peptide to lysine or aspartic acid showed little or no effect on substrate affinity, replacement by arginine increased the Km 35-fold. Similarly, the high Km value of the EGFR pY peptide (Km of 104 microM) derives largely from the arginine residue at the +2 position of the peptide, since arginine to alanine single mutation at the -2 position of the EGFR peptide decreased the Km value 34-fold to 3 microM. Three thiophosphotyrosyl peptides have been prepared and act as substrates and competitive inhibitors of these PTPase catalytic domains.  相似文献   

6.
Leukotriene-A4 hydrolase (EC 3.3.2.6) cleaved the NH2-terminal amino acid from several tripeptides, typified by arginyl-glycyl-aspartic acid, arginyl-glycyl-glycine, and arginyl-histidyl-phenylalanine, with catalytic efficiencies (kcat/Km) > or = 1 x 10(6) M-1 s-1. This exceeds by 10-fold the kcat/Km for its lipid substrate leukotriene A4. Catalytic efficiency declined for dipeptides which had kcat/Km ratios 10-100-fold lower than tripeptides. Tetrapeptides and pentapeptides were even poorer substrates with catalytic efficiencies below 10(3) M-1 s-1. The enzyme preferentially hydrolyzed tripeptide substrates and single amino acid p-nitroanilides with L-arginine at the NH2 terminus. Peptides with proline at the second position were not hydrolyzed, suggesting a requirement for an N-hydrogen at the peptide bond cleaved. Peptides with a blocked NH2 terminus were not hydrolyzed. The specificity constant (kcat/Km) was optimal at pH 7.2 with pK values at 6.8 and 7.9; binding was maximal at pH 8.0. Serum albumins activated the peptidase, increasing tripeptide affinities (Km) by 3-10-fold and specificities (kcat/Km) by 4-13-fold. Two known inhibitors of arginine peptidases, arphamenine A and B, inhibited hydrolysis of L-arginine p-nitroanilide with dissociation constants = 2.0 and 2.5 microM, respectively. Although the primary role of LTA4 hydrolase is widely regarded as the conversion of the lipid substrate leukotriene A4 into the inflammatory lipid mediator leukotriene B4, our data are the first showing that tripeptides are "better" substrates. This is compatible with a biological role for the peptidase activity of the enzyme and may be relevant to the distribution of the enzyme in organs like the ileum, liver, lung, and brain. We present a model which accommodates the available data on the interaction of substrates and inhibitors with the enzyme. This model can account for overlap in the active site for hydrolysis of leukotriene A4 and peptide or p-nitroanilide substrates.  相似文献   

7.
The zymogen and mature enzyme forms of trypsin-like serine proteases exhibit a wide range of activities. The prototypical trypsinogen-trypsin system is an example of a minimally active zymogen and a maximally active mature protease. The present work identifies several features of trypsinogen which govern its activity. Our results indicate that rat trypsin is 10(8)-fold more active than rat trypsinogen. Rat trypsinogen appears to be less active than bovine trypsinogen. His40 is believed to be an important determinant of zymogen activity. We are unable to verify this role for His40 in trypsinogen since the mutation of His40 to Phe appears to change the trypsin-substrate interface. Deletion of the N-terminal Ile16 from trypsin is expected to produce a trypsinogen-like protein since the Ile16-Asp194 salt bridge cannot form. Such mutants have higher activity and BPTI affinity than trypsinogen, which indicates that the activation peptide stabilizes the inactive trypsinogen conformation. The mutation of Lys15 to Ala increases the BPTI affinity and activity of trypsinogen to an even greater extent; thus, removal of Lys15 can account for the effect of the loss of the activation peptide. These results suggest that Lys15 is an important determinant of zymogen activity. The mutation of Asp194 to Asn also increases the BPTI affinity and activity of trypsinogen. This result suggests that in addition to stabilizing the active conformation of trypsin via the Ile16-Asp194 salt bridge, Asp194 also maintains the inactive conformation of trypsinogen. A correlation exists between the values of kcat/Km and BPTI affinity of mutant trypsinogens and trypsins. However, the slope of this correlation is 0.64, which indicates that different "active" conformations are involved in BPTI binding and substrate hydrolysis. DeltaI16V17 trypsinogen is the lone outlier; its BPTI affinity is higher than would be expected based on the value of kcat/Km. We show that the rate of BPTI association is slower for DeltaI16V17 trypsinogen than for a mutant trypsinogen with a similar BPTI affinity. This observation suggests that BPTI binds to an "active" trypsinogen conformation that is not kinetically accessible to substrates.  相似文献   

8.
Trypsinogen is converted to trypsin by the removal of a peptide from the N terminus, which permits formation of a salt bridge between the new N-terminal Ile (residue 16) and Asp194. Formation of this salt bridge triggers a conformational change in the "activation domain" of trypsin, creating the S1 binding site and oxyanion hole. Thus, the activation of trypsinogen appears to represent an example of protein folding driven by electrostatic interactions. The following trypsin mutants have been constructed to explore this problem: Asp194Asn, Ile16Val, Ile16Ala, and Ile16Gly. The bovine pancreatic trypsin inhibitor (BPTI), benzamidine, and leupeptin affinities and activity and pH-rate profiles of these mutants have been measured. The changes in BPTI and benzamidine affinity measure destabilization of the activation domain. These experiments indicate that hydrophobic interactions of the Ile16 side chain provide 5 kcal/mol of stabilization energy to the activation domain while the salt bridge accounts for 3 kcal/mol. Thus, hydrophobic interactions provide the majority of stabilization energy for the trypsinogen to trypsin conversion. The pH-rate profiles of I16A and I16G are significantly different than the pH-rate profile of trypsin, further confirming that the activation domain has been destabilized. Moreover, these mutations decrease kcat/Km and leupeptin affinity in parallel with the decrease in stability of the activation domain. Acylation is selectively decreased, while substrate binding and deacylation are not affected. Together these observations indicate that the stability of protein structure is an important component of transition state stabilization in enzyme catalysis. These results also suggest that active zymogens can be created without providing a counterion for Asp194, and thus have important implications for the elucidation of the structural features which account for the zymogen activity of tissue plasminogen activator and urokinase.  相似文献   

9.
The nonstructural protein NS3 of the hepatitis C virus (HCV) harbors a serine protease domain that is responsible for most of the processing events of the nonstructural region of the polyprotein. Its inhibition is presently regarded as a promising strategy for coping with the disease caused by HCV. In this work, we show that the NS3 protease undergoes inhibition by the N-terminal cleavage products of substrate peptides corresponding to the NS4A-NS4B, NS4B-NS5A, and NS5A-NS5B cleavage sites, whereas no inhibition is observed with a cleavage product of the intramolecular NS3-NS4A junction. The Ki values of the hexamer inhibitory products [Ki(NS4A) = 0.6 microM, Ki(NS5A) = 1.4 microM, and Ki(NS4B) = 180 microM] are lower than the Km values of the respective substrate peptides [Km(NS4A-NS4B) = 10 microM, Km(NS5A-NS5B) = 3.8 microM, and Km(NS4B-NS5A) > 1000 microM]. Mutagenesis experiments have identified Lys136 as an important determinant for product binding. The phenomenon of product inhibition can be exploited to optimize peptide inhibitors of NS3 protease activity that may be useful in drug development.  相似文献   

10.
An antibody generated to an alpha-keto amide containing hapten 1 catalyzes the cis-trans isomerization of peptidyl-prolyl amide bonds in peptides and in the protein RNase T1. The antibody-catalyzed peptide isomerization reaction showed saturation kinetics for the cis-substrate, Suc-Ala-Ala-Pro-Phe-pNA, with a kcat/Km value of 883 s-1.M-1; the reaction was inhibited by the hapten analog 13 (Ki = 3. 0 +/- 0.4 microM). Refolding of denatured RNase T1 to its native conformation also was catalyzed by the antibody, with the antibody-catalyzed folding reaction inhibitable both by the hapten 1 and hapten analog 13. These results demonstrate that antibodies can catalyze conformational changes in protein structure, a transformation involved in many cellular processes.  相似文献   

11.
Quantitative characterization of the interaction of des-kringle1-5-plasmin (microplasmin) with fibrin(ogen) and plasma protease inhibitors may serve as a tool for further evaluation of the role of kringle domains in the regulation of fibrinolysis. Comparison of fibrin(ogen) degradation products yielded by plasmin, miniplasmin (des-kringle1-4-plasmin), microplasmin, and trypsin on SDS gel electrophoresis indicates that the differences in the enzyme structure result in different rates of product formation, whereas the products of the four proteases are very similar in molecular weight. Kinetic parameters show that plasmin is the most efficient enzyme in fibrinogen degradation, and the kcat/KM ratio decreases in parallel with the loss of the kringle domains. The catalytic sites of the four proteases have similar affinities for fibrin (KM values between 0.12 and 0.21 microM). Trypsin has the highest catalytic constant for fibrin digestion (kcat = 0.47 s-1), and among plasmins with different kringle structures, the loss of kringle5 results in a markedly lower catalytic rate constant (kcat = 0.0076 s-1 for microplasmin vs 0.048 s-1 for miniplasmin and 0.064 s-1 for plasmin). In addition, microplasmin is inactivated by plasmin inhibitor (k" = 3.9 x 10(5) M-1 s-1) and antithrombin (k" = 1.4 x 10(3) M-1 s-1) and the rate of inactivation decreases in the presence of fibrin(ogen). Heparin (250 nM) accelerates the inactivation of microplasmin by antithrombin (k" = 10.5 x 10(3) M-1 s-1 ), whereas that by plasmin inhibitor is not affected (k" = 4.2 x 10(5) M-1 s-1).  相似文献   

12.
Acrolein, a highly cytotoxic aldehyde, is a metabolic by-product of the antineoplastic agent cyclophosphamide and is responsible for the development of hemorrhagic cystitis, a serious side effect of cyclophosphamide therapy. Aldose reductase (EC 1.1.1.21), a member of the aldo-keto reductase superfamily, catalyzes the NADPH-dependent reduction of acrolein to allyl alcohol (Km = 80 microM, kcat = 87 min-1). Aldose reductase is expressed at different levels in individuals. This suggests that individual differences in the reductive metabolism of acrolein may be a determinant of acrolein toxicity. In addition to being a substrate, acrolein also produces a time-dependent 7-20-fold increase in the activity of aldose reductase toward a variety of substrates. This involves initial binding of acrolein to a second site (Ks = 58 microM). Acrolein activation of aldose reductase results not only in higher kcat values for all substrates but also in higher Km values and decreased catalytic efficiencies. Acrolein activation of aldose reductase reduces its affinity for aldose reductase inhibitors.  相似文献   

13.
The biologically active isomer of 5-fluoro-5,6-dihydrouracil [(R)-5-fluoro-5,6-dihydrouracil, R-FUH2] was synthesized to study the kinetics of its enzymatic oxidation and hydrolysis by homogeneous dihydropyrimidine dehydrogenase (DPDase) and dihydropyrimidine aminohydrolase (DPHase), respectively. DPDase catalyzed the slow oxidation of R-FUH2 at pH 8 and 37 degrees with a Km of 210 microM and a kcat of 0.026 sec-1 at a saturating concentration of NADP+. The catalytic efficiency (kcat/Km) of DPDase for R-FUH2 was 1/14th of that for 5,6-dihydrouracil (UH2). In the opposite direction, DPDase catalyzed the reduction of 5-fluorouracil (FU) with a Km of 0.70 microM and a kcat of 3 sec-1 at a saturating concentration of NADPH. Thus, DPDase catalyzed the reduction of FU 30,000-fold more efficiently than the oxidation of R-FUH2. In contrast to the slow oxidation of R-FUH2 by DPDase, R-FUH2 was hydrolyzed very efficiently by DPHase with a Km of 130 microM and a kcat of 126 sec-1. The catalytic efficiency of DPHase for the hydrolysis of R-FUH2 was approximately twice that for the hydrolysis of UH2. Because R-FUH2 is hydrolysis of R-FUH2 was approximately twice that for the hydrolysis of UH2. Because R-FUH2 is hydrolyzed considerably more efficiently than it is oxidized and because the activity of DPHase was 250- to 500-fold greater than that of DPDase in bovine and rat liver, the hydrolytic pathway should predominate in vivo.  相似文献   

14.
Mammalian alkaline phosphatases (APs) display 10-100-fold higher kcat values than do bacterial APs. To begin uncovering the critical residues that determine the catalytic efficiency of mammalian APs, we have compared the sequence of two bovine intestinal APs, i.e. a moderately active isozyme (bovine intestinal alkaline phosphatase, bIAP I, approximately 3,000 units/mg) previously cloned in our laboratory, and a highly active isozyme (bIAP II, approximately 8, 000 units/mg) of hitherto unknown sequence. An unprecedented level of complexity was revealed for the bovine AP family of genes during our attempts to clone the bIAP II cDNA from cow intestinal RNAs. We cloned and characterized two novel full-length IAP cDNAs (bIAP III and bIAP IV) and obtained partial sequences for three other IAP cDNAs (bIAP V, VI, and VII). Moreover, we identified and partially cloned a gene coding for a second tissue nonspecific AP (TNAP-2). However, the cDNA for bIAP II, appeared unclonable. The sequence of the entire bIAP II isozyme was determined instead by a classical protein sequencing strategy using trypsin, carboxypeptidase, and endoproteinase Lys-C, Asp-N, and Glu-C digestions, as well as cyanogen bromide cleavage and NH2-terminal sequencing. A chimeric bIAP II cDNA was then constructed by ligating wild-type and mutagenized fragments of bIAP I, III, and IV to build a cDNA encoding the identified bIAP II sequence. Expression and enzymatic characterization of the recombinant bIAP I, II, III, and IV isozymes revealed average kcat values of 1800, 5900, 4200, and 6100 s-1, respectively. Comparison of the bIAP I and bIAP II sequences identified 24 amino acid positions as likely candidates to explain differences in kcat. Site-directed mutagenesis and kinetic studies revealed that a G322D mutation in bIAP II reduced its kcat to 1300 s-1, while the converse mutation, i.e. D322G, in bIAP I increased its kcat to 5800 s-1. Other mutations in bIAP II had no effect on its kinetic properties. Our data clearly indicate that residue 322 is the major determinant of the high catalytic turnover in bovine IAPs. This residue is not directly involved in the mechanism of catalysis but is spatially sufficiently close to the active site to influence substrate positioning and hydrolysis of the phosphoenzyme complex.  相似文献   

15.
An extracellular beta-glucosidase from Fusarium oxysporum was purified to homogeneity by gel-filtration and ion-exchange chromatographies. The enzyme, a monomeric protein of 110 kDa, was maximally active at pH 5.0-6.0 and at 60 degrees C. It hydrolysed 1-->4-linked aryl-beta-glucosides and 1-->4-linked, 1-->3-linked and 1-->6-linked beta-glucosides. The apparent Km and kcat values for p-nitrophenyl beta-D-glucopyranoside (4-NpGlcp) and cellobiose were 0.093 (Km), 1.07 mM (kcat) and 1802 (Km), 461.5 min-1 (kcat), respectively. Glucose and gluconolactone inhibited the enzyme competitively with Ki values of 2.05 mM and 3.03 microM, respectively. Alcohols activated the enzyme; butanol showed maximum effect (2.2-fold at 0.5 M) while methanol increased the activity by 1.4-fold at 1 M. The enzyme catalysed the synthesis of methylglucosides, ethylglucoside and propylglucosides, as well as trisaccharides in the presence of different alcohols and disaccharides, respectively. In addition, the enzyme hydrolysed the unsubstituted and methylumbelliferyl cello-oligosaccharides [MeUmb(Glc)n] but the rate of hydrolysis decreased with increasing chain length. Analysis of products released from MeUmb(Glc)n as a function of time revealed that the enzyme attacked these substrates in a stepwise manner and from both ends. Thus, beta-glucosidase from F. oxysporum, with the above interesting properties, could be of commercial interest.  相似文献   

16.
The gene for chorismate mutase (CM) from the archaeon Methanococcus jannaschii, an extreme thermophile, was subcloned and expressed in Escherichia coli. This gene, which belongs to the aroQ class of CMs, encodes a monofunctional enzyme (AroQf) able to complement the CM deficiency of an E. coli mutant strain. The purified protein follows Michaelis-Menten kinetics (kcat = 5.7 s-1 and Km = 41 microM at 30 degreesC) and displays pH-independent activity in the range of pH 5-9. Its activation parameters [Delta H = 16.2 kcal/mol, Delta S = -1. 7 cal/(mol.K)] are similar to those of another well characterized AroQ class CM, the mesophilic AroQp domain from E. coli. Like AroQp, the thermophilic CM is an alpha-helical dimer, but approximately 5 kcal/mol more stable than its mesophilic counterpart as judged from equilibrium denaturation studies. The possible origins of the thermostability of M. jannaschii AroQf, the smallest natural CM characterized to date, are discussed in light of available sequence and tertiary structural information.  相似文献   

17.
Duodenase, a serine protease from bovine duodenum mucosa, was located in endoplasmic reticulum, the Golgi secretory granules of epithelial cells and ducts of Brunner's glands by the A-gold immunocytochemical method. Duodenase exhibits trypsin-like and chymotrypsin-like specificities with a preference for substrates having lysine at the P1 and proline at the P2 positions. The kinetic constants for the hydrolysis of 21 potential duodenase substrates are reported. The best substrates were found to be alpha-N-tosylglycylprolyllysine 4-nitroanilide (k[cat]/Km of 35000 M[-1] s[-1]), alpha-N-succinylthreonylprolyllysine 4-nitroanilide (k[cat]/Km of 18000 M[-1] s[-1]) and alpha-N-serylprolyllysine 4-nitroanilide (k[cat]/Km of 2600 m[-1] s[-1]), all of which contain the P1-P3 sequence of the enteropeptidase zymogen/activation site. On the basis of its catalytic properties and sites of localization, duodenase has been postulated to be an activator of the enteropeptidase precursor. A tetradecapeptide (LVTQEVSPKIVGGS) having the P9-P5'sequence of the cleavage site of zymogen activation of bovine proenteropeptidase was synthesized, and kinetic parameters of its hydrolysis by duodenase were determined (Km of 87 microM; k[cat] of 1.4 s[-1]; k[cat]/Km of 16000 M[-1] s[-1]). Crystals of duodenase frozen in a stream of liquid nitrogen diffracted synchrotron X-rays to 0.2-nm resolution.  相似文献   

18.
Calcium is required for the substrate binding and for the chemical step of the interfacial catalytic turnover cycle of pancreatic phospholipase A2 (PLA2), but not for the binding of the enzyme to the interface. The role of calcium and other divalent cations (C) is analyzed for the effect on the substrate binding and kcat* for the chemical step. The cofactor role of 3d-cations(II) (C) for the hydrolysis of dimyristoylphosphatidylmethanol (DMPM) vesicles is characterized as an equilibrium dissociation constant for the interfacial binary (E*C) and ternary (E*CL) complexes of PLA2 and substrate mimics (L). Of the cations(II) that promote the binding of a mimic to the enzyme at the interface (E*), only a subgroup supports the chemical step. For example, Cd, Zn, and Cu form ternary E*CL complexes with kcat* of <1 s-1, compared to the rate of >100 s-1 with Ca, Fe, Mn, Co, and Ni. Oxygen exchange from H218O to the products of hydrolysis of DMPM incorporates one 18O in myristate. Incorporation of the first and second 18O occurs during the incubation of both the products of hydrolysis in H218O with PLA2 and Ca, but not with Zn. The cation-dependent changes in the UV difference spectrum, associated with the formation of E*C and E*CL, suggest that the changes are mainly due to catalytic His-48, and possibly Tyr-52 and Tyr-73, and are different with Ca as opposed to Zn. These results and simulations suggest considerable plasticity in the calcium binding and catalytic site environment. It is proposed that the higher ground state stability of the E*CS complex with the inhibitory cations increases the effective activation energy. For the chemical step, calcium coordinated with a nucleophilic water and the ester carbonyl oxygen facilitates the near-attack geometry in the E*CaS, and the His-48.Asp-99 pair acts as a proton acceptor. As a prelude to establishing the catalytic mechanism, factors controlling the energetically demanding transition state are also discussed.  相似文献   

19.
Isopeptidase T (IPaseT) can hydrolyze isopeptide bonds of polyubiquitin (polyUb) chains, simple C-terminal derivatives of Ub, and certain peptides. We recently reported that IPaseT is regulated by ubiquitin (Ub); while submicromolar Ub activates, higher concentrations inhibit this enzyme [Stein et al. (1995) Biochemistry 34, 12616]. To explain these observations, we proposed a model for IPaseT involving two binding sites for Ub. According to the model, the two sites are adjacent to one another and are the extended active site that binds two Ub moieties of a polyUb chain. The "activation site" binds the Ub that donates Lys to the isopeptide bond. The "inhibition site" is adjacent and binds the Ub that donates the C-terminal Gly to the isopeptide bond. We now report that the interaction of IPaseT with the C-terminal aldehyde of Ub (Ub-H) is also modulated by Ub. In the absence of Ub, Ub-H inhibits IPaseT with a Ki of 2.3 nM, while at 0.6 microM Ub, where the "activation site" is occupied, Ki is less than 0.1 nM. At high Ub concentrations, where both the "activation" and "inhibition" sites are occupied, IPaseT cannot bind Ub-H. We also determined the kinetics of inhibition of IPaseT by Ub-H. In the absence of Ub, a two-step mechanism is followed. In the first step, Ub-H slowly combines with IPaseT to form a relatively weak complex (K1 = 260 nM) that slowly isomerizes to the final, stable complex that accumulates in the steady-state (k2 = 2 x 10(-3) s-1; k-2 = 0.02 x 10(-3) s-1). In contrast, Ub-activated IPaseT is inhibited by Ub-H through a three-step process. In the first step, Ub-H rapidly combines with IPaseT to form a complex (K1 = 10 nM) that slowly isomerizes to a second, more stable complex (k2 = 18 x 10(-3) s-1; k-2 = 1.5 x 10(-3) s-1). In the third step, the second complex converts to the final complex (k3 = 1.5 x 10(-3) s-1; k-3 < 0.2 x 10(-3) s-1). To unify the results of this study with our previous results on catalysis, we propose that binding of Ub either to catalytic transition states or to tetrahedral inhibition intermediates liberates more free energy than binding of Ub to the reactant state of IPaseT and that IPaseT can utilize this binding energy to stabilize both of these tetrahedral species. The overall effect is a Ub-induced increase in catalytic efficiency or inhibitory potency.  相似文献   

20.
The phosphotriesterase from Pseudomonas diminuta catalyzes the hydrolysis of a wide array of phosphotriesters and related phosphonates, including organophosphate pesticides and military nerve agents. It has now been shown that this enzyme can also catalyze the hydrolysis of phosphodiesters, albeit at a greatly reduced rate. However, the enzymatic hydrolysis of ethyl-4-nitrophenyl phosphate (compound I) by the wild-type enzyme was >10(8) times faster than the uncatalyzed reaction (kcat = 0.06 s-1 and Km = 38 mM). Upon the addition of various alkylamines to the reaction mixture, the kcat/Km for the phosphodiester (compound I) increased up to 200-fold. Four mutant enzymes of the phosphotriesterase were constructed in a preliminary attempt to improve phosphodiester hydrolysis activity of the native enzyme. Met-317, which is thought to reside in close proximity to the pro-S-ethoxy arm of the paraoxon substrate, was mutated to arginine, alanine, histidine, and lysine. These mutant enzymes showed slight improvements in the catalytic hydrolysis of organophosphate diesters. The M317K mutant enzyme displayed the most improvement in catalytic activity (kcat = 0.34 s-1 and Km = 30 mM). The M317A mutant enzyme catalyzed the hydrolysis of the phosphodiester (compound I) in the presence of alkylamines up to 200 times faster than the wild-type enzyme in the absence of added amines. The neutralization of the negative charge on the oxygen atom of the phosphodiester by the ammonium cation within the active site is thought to be responsible for the rate enhancement by these amines in the hydrolytic reaction. These results demonstrate that an active site optimized for the hydrolysis of organophosphate triesters can be made to catalyze the hydrolysis of organophosphate diesters.  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