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 31 毫秒
1.
The reversible binding constant (Ki) for tetramethylammonium ion (TMA) was determined from the decrease in the bimolecular rate constant (ki) observed with each of 21 organophosphate or carbamate inhibitors of acetylcholinesterase (EC 3.1.1.7). The Ki values obtained were reasonably constant (5.8 X 10(-4) +/- 0.38 M), and this is consistent with reports indicating that TMA binds to a single site on the enzyme.  相似文献   

2.
To evaluate how two inhibitors influence oxidative drug metabolism, this study investigated the inhibitory effects of mexiletine with cimetidine and mexiletine with lidocaine, both individually and in combination, on the oxidative metabolism of two probe substrates, aminopyrine and aniline in rat liver microsomes. Mexiletine was a competitive inhibitor of aminopyrine N-demethylation, whereas cimetidine was a mixed type of inhibitor (Ki = 2.00 +/- 0.04 and 0.20 +/- 0.02 mM, respectively). For aniline hydroxylation, mexiletine exhibited a mixed type of inhibition, whereas lidocaine was a noncompetitive inhibitor (Ki = 0.60 +/- 0.07 and 8.50 +/- 0.12 mM, respectively). The combined inhibition of either mexiletine with cimetidine or mexiletine with lidocaine on aminopyrine and aniline metabolism was close to the fully additive effects of the individual compounds when their individual concentrations were below a 2-fold Ki concentration, regardless of the apparent kinetic inhibition type. The combined inhibition was less than fully additive when the individual concentrations were twice the Ki or above. These results demonstrate that, when two inhibitors of oxidative drug metabolism are combined, both the Ki values and the concentrations of inhibitors play important roles in determining the extent of additive inhibition of enzyme activity.  相似文献   

3.
Enantiomers of N-methyl-N,alpha-methylbenzylbutyramide (1), 1-butyl-3-methyl-3'-alpha-methylbenzylurea (2), 1,2,3, 4-tetrahydro-1-naphthyl-N-butylcarbamate (3), 1,1'-bi-2-naphthyl-2, 2'-di-N-butylcarbamate (4), 1, 1'-bi-2-naphthyl-2-ol-2'-N-butylcarbamate (5), and 1, 1'-bi-2-naphthyl-2-butyrate-2'-N-butylcarbamate (6) are inhibitors of porcine pancreatic cholesterol esterase-catalyzed hydrolysis of 4-nitrophenyl butyrate and of electric eel acetylcholinesterase-catalyzed hydrolysis of acetylthiocholine in the presence of 5,5'-dithiobis-2-nitrobenzoate. For competitive inhibitors, values of the inhibition constant (Ki) and the enantiomeric ratio (Ecomp.) are investigated. For active site-directed irreversible inhibitors, values of the inhibition constant (Ki), the carbamylation constant (k2), the bimolecular rate constant (ki), and the enantiomeric ratio (E) are investigated. Toward both enzymes, compounds 1 are poor competitive inhibitors (Ki=102-104 microM) but have good enantioselectivities (Ecomp.=10-50, the preference for R). R-2 and S-2 are competitive inhibitors of acetylcholinesterase with Ki=26 and 80 microM, respectively (the preference for R) but are active site-directed irreversible inhibitors of cholesterol esterase with ki=4 and 16 M-1 sec-1, respectively (the preference for S). For those competitive inhibitions, both leaving group hydrophilic and hydrophobic binding sites of cholesterol esterase or both anionic substrate binding site and peripheral anionic binding site of acetylcholinesterase bind to N,N-methyl-alpha-methylbenzyl disubstituted amide parts of these inhibitors and the enzyme does not catalyze the hydrolysis of these inhibitors. The opposite stereopreference (S) for the inhibition of cholesterol esterase by compounds 2 may be due to the fact that N, N-methyl-alpha-methylbenzyl disubstituted amide parts of these inhibitors bind to the alkyl chain binding site of the enzyme. Compounds 3-6 are active site-directed irreversible inhibitors of cholesterol esterase (ki=1-13000 M-1 s-1) and peripheral anionic binding site-directed irreversible inhibitors of acetylcholinesterase (ki=1.7-1300 M-1 s-1). Compounds 3 have low enantioselectivities (E=1.3-1.4) for both enzymes. The stereopreference for atropisomers 4 and 6 is S-form toward both enzymes (E=2-30) and is identical to that of cholesterol esterase-catalyzed hydrolysis of 1,1'-bi-2-naphthyl-2,2'-diacylate. This stereopreference (S) may be due to the fact that the butyryl group or one of two butylcarbamate groups of S-atropisomers binds more effectively to the leaving group hydrophobic binding site of cholesterol esterase or the peripheral anionic binding site of acetylcholinesterase than that of R-atropisomers. The opposite stereopreference (R) for atropisomers 5 toward both enzymes may be due to a favorable interaction between the hydroxyl group of the inhibitors and the leaving group hydrophilic binding site of cholesterol esterase or the peripheral anionic binding site of acetylcholinesterase.  相似文献   

4.
The binding of two different reaction products (p-nitrobenzyl glutathione and the aflatoxin-glutathione conjugate) to mouse glutathione S-transferase A3-3 (mGSTA3-3) has been measured using equilibrium dialysis and a direct fluorescence quenching technique. As expected, p-nitrobenzyl glutathione was found to bind with a stoichiometry of 2.24 +/- 0.17 mol/mol of dimeric enzyme. However, the much larger aflatoxin-glutathione conjugate, 8, 9-dihydro-8-(S-glutathionyl)-9-hydroxyl-aflatoxin B1 (AFB-GSH), was found to bind with a stoichiometry of 1.12 +/- 0.08 mol/mol of dimeric enzyme. p-Nitrobenzyl glutathione bound mGSTA3-3 with a dissociation constant (Kd) of 59 +/- 17 microM while the aflatoxin-glutathione conjugate bound the enzyme with a Kd of 0.86 +/- 0.19 microM. Glutathione competitively inhibited binding of AFB-GSH to mGSTA3-3 with a Ki of 1.5 mM, suggesting that AFB-GSH was binding to the enzyme active site. Although AFB-GSH bound to mGSTA3-3 with a stoichiometry of 1 mol/mol of dimeric enzyme, AFB-GSH completely inhibited activity toward 1-chloro-2, 4-dinitrobenzene, indicating that AFB-GSH binding to one active site alters affinity for 1-chloro-2,4-dinitrobenzene in the active site of the other subunit. To our knowledge, this is the first report of a glutathione S-transferase reaction product which binds to the enzyme with a stoichiometry of 1 mol/mol of dimer.  相似文献   

5.
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   

6.
Woodward's reagent K (WRK) completely inactivated Escherichia coli uridine phosphorylase by reversible binding in the active site (Ki = 0.07 mM) with subsequent modification of a carboxyl (k2 = 1.2 min-1). Neither substrate alone protected uridine phosphorylase from inactivation. The presence of phosphate did not affect the Ki and k2 values. The addition of uracil or uridine led to a significant increase of both Ki (to 2.5 or 2.1 mM, respectively) and k2 (to 6.1 or 4.8 min-1, respectively) values. Thus, WRK could react in accordance with slow (high affinity) and fast (low affinity) mechanisms. Combined addition of phosphate and uracil completely protected uridine phosphorylase. Tryptic digestion yielded a single modified peptide (Ser4-Asp(WRK)-Val-Phe-His-Leu-Gly-Leu-Thr-Lys13). Treatment of the modified enzyme with hydroxylamine led to removal of the bulky WRK residue and replacement of the Asp5 carboxyl by a hydroxamic group. The enzyme thus obtained recovered about 10% of initial specific activity, whereas its substrate binding ability changed only moderately; the Km values for phosphate and uridine were changed from 5.1 and 0.19 mM (or 7.3 and 0.14 mM according to Leer et al. (Leer, J.C., Hammer-Jespersen, K., and M. Schwartz (1977) Eur. J. Biochem. 75, 217-224)) to 22.6 and 0.12 mM, respectively. The hydroxamic enzyme had higher thermostability than the native enzyme. The results obtained demonstrated the importance of the carboxyl at position 5. The loss of activity after selective group replacement is due to impaired stabilization of the transition state rather than to a decline in substrate affinity or change of the active site structure.  相似文献   

7.
Synaptosomes prepared from mouse brain possess a Na+-dependent transport system for gamma-hydroxybutyrate displaying saturation kinetics, the transport constant (Kt) for which was calculated as 31 +/- 9 micromol/l. Several gamma-hydroxybutyrate and gamma-aminobutyric acid (GABA) structural analogues were tested as potential inhibitors of gamma-hydroxybutyrate transport. The most effective inhibitor was harmaline (Ki = 94 +/- 21 micromol/l), a known competitive inhibitor of Na+ binding to certain transport proteins. 2-Hydroxycinnamic acid, 3-(2-furyl)acrylic acid and citrazinic acid also inhibited transport and were competitive with respect to gamma-hydroxybutyrate. The least effective gamma-hydroxybutyrate analogues were 3-hydroxypropane sulfonic acid (Ki = 4.1 +/- 0.8 mmol/l) 3,5-dihydroxybenzoic acid (Ki = 6.1 +/- 2. 8 mmol/l) and 3-hydroxybenzoic acid (Ki = 6.9 +/- 3.3 mmol/l), although 2-hydroxypropane sulfonic acid and kynurenic acid had no measurable effects. Four inhibitors of GABA transport - nipecotic acid, guvacine, ketamine and beta-alanine and GABA itself, were without effect on gamma-hydroxybutyrate transport. These results show that certain drugs that structurally resemble gamma-hydroxybutyrate have the capacity to compete with gamma-hydroxybutyrate at its recognition site on the transporter. By examining the structure of such inhibitors, we can learn more about the properties of the substrate binding site on the carrier protein. Moreover, the absence of inhibition by GABA uptake inhibitors shows that gamma-hydroxybutyrate transport is a separate entity from GABA transport.  相似文献   

8.
An enzyme with alpha-L-rhamnosidase activity was purified by anion exchange chromatography from an Aspergillus niger commercial preparation. The alpha-L-rhamnosidase was shown to be N-glycosylated, and had a molecular mass of 85 kD on sodium dodecylsulfate-polyacrylamide gel electrophoresis of which approximately 12% was contributed by carbohydrate. The enzyme was optimally active at pH 4.5 and 65 degrees C. When tested towards p-nitrophenyl-alpha-L-rhamnopyranoside it showed Km and Vmax values of 2.9 mM and 20.6 U mg-1, respectively whereas it was inhibited competitively by L-rhamnose (Ki 3.5 mM). Substrate specificity studies showed alpha-L-rhamnosidase to be active both on alpha-1,2 and alpha-1,6 linkages to beta-D-glucose. Moreover, the enzyme was able to release L-rhamnose from geranyl-beta-D-rutinoside and 2-phenylethyl-beta-D-rutinoside.  相似文献   

9.
Resistance of HIV-1 to protease inhibitors has been associated with changes at residues Val82 and Ile84 of HIV-1 protease (HIV PR). Using both an enzyme assay with a peptide substrate and a cell-based infectivity assay, we examined the correlation between the inhibition constants for enzyme activity (Ki values) and viral replication (IC90 values) for 5 active site mutants and 19 protease inhibitors. Four of the five mutations studied (V82F, V82A, I84V, and V82F/I84V) had been identified as conferring resistance during in vitro selection using a protease inhibitor. The mutant protease genes were expressed in Escherichia coli for preparation of enzyme, and inserted into the HXB2 strain of HIV for test of antiviral activity. The inhibitors included saquinavir, indinavir, nelfinavir, 141W94, ritonavir (all in clinical use), and 14 cyclic ureas with a constant core structure and varying P2, P2' and P3, P3' groups. The single mutations V82F and I84V caused changes with various inhibitors ranging from 0.3- to 86-fold in Ki and from 0.1- to 11-fold in IC90. Much larger changes compared to wild type were observed for the double mutation V82F/I84V both for Ki (10-2000-fold) and for IC90 (0.7-377-fold). However, there were low correlations (r2 = 0.017-0.53) between the mutant/wild-type ratio of Ki values (enzyme resistance) and the mutant/wild-type ratio of viral IC90 values (antiviral resistance) for each of the HIV proteases and the viruses containing the identical enzyme. Assessing enzyme resistance by "vitality values", which adjust the Ki values with the catalytic efficiencies (kcat/Km), caused no significant improvement in the correlation with antiviral resistance. Therefore, our data suggest that measurements of enzyme inhibition with mutant proteases may be poorly predictive of the antiviral effect in resistant viruses even when mutations are restricted to the protease gene.  相似文献   

10.
Glucuronidation of drugs represents a major pathway of human drug metabolism. Numerous studies show that the glucuronides formed can accumulate during chronic therapy and/or have direct pharmacological activity. In both cases, cleavage of the glucuronide by human beta-glucuronidase (beta-Gluc) would release the parent compound, thereby modifying drug disposition. Variability in expression of beta-Gluc could therefore be a confounding factor for interindividual variability in drug disposition both in the setting of accumulating glucuronides or for the use of glucuronides as prodrugs, such as the nontoxic glucuronide-spacer derivative of doxorubicin (Dox-S-G). We therefore investigated expression and function of beta-Gluc in human liver (n = 30) and human kidney (n = 18). Cleavage of the model compound 4-methylumbelliferyl-beta-D-glucuronide (MUG) revealed a wide range of activities in liver (0.32-1.85 mumol/mg/h, mean value 0.87 +/- 0.34 mumol/mg/h) and kidney (0.07-1.00 mumol/mg/h, mean 0.39 +/- 0.21 mumol/mg/h), which followed a log normal distribution. Variable enzyme activity was closely correlated to enzyme expression as assessed by Western blotting (r = 0.80, P < .001 and r = 0.71, P < .05 for liver and kidney, respectively). Glycyrrhizin (Ki = 470 and 570 microM), estradiol 3-glucuronide (Ki = 0.9 and 1.2 mM) and paracetamol glucuronide (Ki = 1.6 and 2 mM) were found to inhibit beta-Gluc activity competitively in liver and kidney, respectively. Enzyme kinetics were investigated in detail for MUG and Dox-S-G. Whereas MUG followed monophasic Michaelis-Menten kinetics in liver (K(m) = 1.32 +/- 0.25 mM, Vmax = 1201 +/- 462 nmol/mg/h, n = 3) and kidney (K(m) = 1.04 +/- 0.05 mM, Vmax = 521 +/- 267 nmol/mg/h, n = 3), cleavage of Dox-S-G was best described by the Hill equation, which indicated a cooperative substrate binding pattern of Dox-S-G. In summary, beta-Gluc function shows wide interindividual variability in human liver and kidney that is due to different steady-state levels of the enzyme. Moreover, enzyme kinetics are substrate-dependent, with Dox-S-G showing a cooperative binding. These data indicate the possibility of wide interindividual variability in beta-Gluc-mediated cleavage of drug glucuronides in the human.  相似文献   

11.
Cytosolic and lysosomal acid phosphatases have the ability to hydrolyze orthophosphoric monoesters below pH 5-6. However, it is thought they may have different intracellular roles. To clarify their properties, substrate specificity, inhibitor sensitivity and the modulation of enzyme by redox conditions were determined using bovine liver enzymes. DEAE-cellulose chromatography following (NH4)2SO4 fractionation revealed three forms of cytosolic acid phosphatases as in the KCl gradient (0-500 mM). After Sephadex G-75 gel filtration, the enzymes appeared as single bands on SDS-polyacrylamide gel electrophoresis (SDS-PAGE). Their activities for D-erythrose 4-phosphate co-purified with p-nitrophenylphosphatase activities in all steps. In contrast the lysosomal enzyme was purified by Octyl-Sepharose column chromatography after n-butanol treatment, (NH4)2SO4 fractionation, Bio gel P-200 gel filtration and DE-52 chromatography. The relative molecular masses (M(r)) determined by SDS-PAGE indicated that M(r) of the cytosolic enzymes (16,000) was less that of lysosomal enzyme (160,000). The cytosolic enzymes were active against sugar phosphates and were inhibited by 1 mM Cu2+. In addition, the cytosolic enzymes were inactivated by 5 mM oxidized glutathione and protected by 10 mM reduced glutathione (in the presence or absence of thioltransferase), suggesting that sensitive cysteinyl residue(s) existed. The lysosomal enzyme was active against various substrates and was strongly inhibited by 1 mM Cu2+ and 2 mM fluoride. The results presented here suggest that cytosolic enzymes have different properties from those of lysosomal enzyme with respect to substrates, inhibitors and regulation of activity.  相似文献   

12.
The properties are discussed of system y+L, a broad scope amino acid transporter which was first identified in human erythrocytes. System y+L exhibits two distinctive properties: (a) it can bind and translocate cationic and neutral amino acids, and (b) its specificity varies depending on the ionic composition of the medium. In Na+ medium, the half-saturation constant for L-lysine influx was 9.5 +/- 0.67 microM and the inhibition constant (Ki) for L-leucine was 10.7 +/- 0.72 microM. L-Leucine is the neutral amino acid that binds more powerfully, whereas smaller analogues, such as L-alanine and L-serine interact less strongly (the corresponding inhibition constants were Ki,Ala, 0.62 +/- 0.11 mM; Ki,Ser, 0.49 +/- 0.08 mM). In the presence of K+, the carrier functions as a cationic amino acid specific carrier, but Li+ is able to substitute for Na+ facilitating neutral amino acid binding. The effect of the inorganic cations is restricted to the recognition of neutral amino acids; translocation occurs at similar rates in the presence of Na+, K+ and Li+. The only structural feature that appears to impair translocation is bulkiness and substrates with half-saturation constants differing by more than 100-fold translocate at the same rate. This suggests that translocation is largely independent of the forces of interaction between the substrate and the carrier site. System y+L activity has been observed in Xenopus laevis oocytes injected with the cRNA for the heavy chain of the 4F2 human surface antigen. 4F2hc is an integral membrane protein with a single putative membrane-spanning domain and it remains to be clarified whether it is part of the transporter or an activator protein.  相似文献   

13.
The interactions between bovine pancreatic ribonuclease A (RNase A) and its RNA substrate extend beyond the scissile P-O5' bond. Enzymic subsites interact with the bases and phosphoryl groups of the bound substrate. Those residues interacting with the phosphoryl group comprise the P0, P1, and P2 subsites, with the scissile bond residing in the P1 subsite. Here, the function of the P0 and P2 subsites of RNase A is characterized in detail. Lys66 (P0 subsite) and Lys7 and Arg10 (P2 subsite) were replaced with alanine residues. Wild-type RNase A and the K66A, K7A/R10A, and K7A/R10A/K66A variants were evaluated as catalysts for the cleavage of poly(cytidylic acid) [poly(C)] and for their abilities to bind to single-stranded DNA, a substrate analogue. The values of kcat and Km for poly(C) cleavage were affected by altering the P0 and P2 subsites. The kcat/Km values for poly(C) cleavage by the K66A, K7A/R10A, and K7A/R10A/K66A variants were 3-fold, 60-fold, and 300-fold lower, respectively, than that of wild-type RNase A. These values indicate that the P0 and P2 subsites contribute 0.70 and 2.46 kcal/mol, respectively, to transition-state binding. Binding experiments indicate that the P0 and P2 subsites contribute 0.92 and 1.21 kcal/mol, respectively, to ground-state binding. Thus, the P0 subsite makes a uniform contribution toward binding the ground state and the transition state, whereas the P2 subsite differentiates, binding more tightly to the transition state than to the ground state. In addition, nucleic acid binding to wild-type RNase A is strongly dependent on NaCl concentration, but this dependence is diminished upon alteration of the P0 or P2 subsite. The logarithm of Kd is a linear function of the logarithm of [Na+] over the range 0.018 M 相似文献   

14.
S 5627 is a synthetic analogue of chlorogenic acid. S 5627 is a potent linear competitive inhibitor of glucose 6-phosphate (Glc-6-P) hydrolysis by intact microsomes (Ki = 41 nM) but is without effect on the enzyme in detergent- or NH4OH-disrupted microsomes. 3H-S 5627 was synthesized and used as a ligand in binding studies directed at characterizing T1, the Glc-6-P transporter. Binding was evaluated using Ca2+-aggregated microsomes, which can be sedimented at low g forces. Aside from a modest reduction in K values for both substrate and S 5627, Ca2+ aggregation had no effect on glucose-6-phosphatase (Glc-6-Pase). Scatchard plots of binding data are readily fit to a simple "two-site" model, with Kd = 21 nM for the high affinity site and Kd = 2 microM for the low affinity site. Binding to the high affinity site was competitively blocked by Glc-6-P (Ki = 9 microM), whereas binding was unaffected by mannose-6-phosphate, Pi, and PPi and only modestly depressed by 2-deoxy-D-glucose 6-phosphate, a poor substrate for Glc-6-Pase in intact microsomes. Thus the high affinity 3H-S 5627 binding site fits the criteria for T1. Permeabilization of the membrane with 0.3% (3-[(chloramidopropyl)-dimethylammonio]-1-propanesulfonate) activated Glc-6-Pase and broadened its substrate specificity, but it did not significantly alter the binding of 3H-S 5627 to the high affinity sites or the ability of Glc-6-P to block binding. These data demonstrate unequivocally that two independent Glc-6-P binding sites are involved in the hydrolysis of Glc-6-P by intact microsomes. The present findings are the strongest and most direct evidence to date against the notion that the substrate specificity and the intrinsic activity of Glc-6-Pase in native membranes are determined by specific conformational constraints imposed on the enzyme protein. These data constitute compelling evidence for the role of T1 in Glc-6-Pase activity.  相似文献   

15.
We have investigated the binding interactions between highly de-N-acetylated chitosans and lysozyme from chicken egg white by one-dimensional and two-dimensional 1H-NMR spectroscopy. A fully de-N-acetylated chitosan (fraction of N-acetylated units, F < 0.001) induced no observable changes in the 1H chemical shifts of lysozyme. However, a chitosan with F(A) = 0.04, where the N-acetylated units are predominantly surrounded by de-N-acetylated units (a monoacetylated sequence), induced significant shifts of several lysozyme resonances, demonstrating a specific interaction between lysozyme and de-N-acetylated units in the chitosan. The interaction between the two positively charged molecules increased with increasing ionic strength, as expected. The dissociation constant (Kd) between lysozyme and the monoacetylated sequence was strongly dependent on pH* (pH measured in D2O), with Kd = 0.02+/-0.01 mM at pH* 6.0, Kd = 0.11+/-0.02 mM at pH* 4.5, and Kd approximately 2 mM at pH* 3, suggesting that electrostatic forces contribute to the observed binding. The complex was strikingly stable, with bound lifetimes in the range of 10-25 ms at pH* 4.5 and 328-300 K. Most lysozyme resonances that were affected by the binding were assigned, and we suggest that the monoacetylated chitosan sequence binds to the active site cleft of lysozyme with the N-acetylated unit in subsite C. Assuming this binding mode, we have discussed the contributions in energetic terms from individual subsites of lysozyme towards binding of N-acetylated and de-N-acetylated units.  相似文献   

16.
The S1'-S3' subsite specificity of prolyl endopeptidase from Flavobacterium meningoseptum was studied by acyl transfer to libraries of amino acid amides and peptides. Whereas the S1' and S3' subsites influence the specificity for the amino component by approximately one order of magnitude, the S2' subsite possesses a markedly higher specificity. Besides the high specificity for hydrophobic residues at P1'-P3', proline was efficiently bound by the S2' and S3' subsites of the enzyme. In contrast, no binding of P1' proline-containing peptides was observed. It could be demonstrated that the specificity of the S' subsite is not restricted to L-amino acids. Effective P'-S' interactions were also found for beta- and gamma-amino acids indicating that the enzyme does not form close contacts to the backbone of P1' and P2' amino acid residues.  相似文献   

17.
Binding of ATP to bovine serum albumin was shown by ultrafiltration and NMR. The binding was pH dependent. Scatchard analysis revealed that at pH 5.4, 6.4 and 7.4, dissociation constant Kd was 13, 40 and 120 microM, respectively, and no binding was observed at pH 8.4. The binding stoichiometry was 1:1 for all pH. Dimer of BSA did not bind ATP. From chemical shifts of 31P-NMR, Kd was estimated to be 15 microM at pH 5.4, which is very close to that determined by ultrafiltration. While adenosine did not interfere with the binding. GTP, dCTP, ADP, UTP, AMP, phosphate and pyrophosphate were competitive inhibitors and their inhibition constants Ki were 25, 32, 36, 50, 130, 1000 and 186 microM, respectively. Fatty acids such as lauric acid and palmitic acid did not interfere with the binding. Warfarin was a non-competitive inhibitor. Cl- competitively inhibited the binding, and the inhibition constant was 20 mM. The dissociation constants of the Cl- binding were reported to be 0.42 mM for the first binding site, 10-5 mM for the second and 303-143 mM for the third [G. Scatchard, W.T. Yap, J. Am. Chem. Soc., 86 (1964) 3434; G. Scatchard et al., J. Am. Chem. Soc. 79 (1957) 12]. This suggests that the ATP binding site may be the second Cl- binding site.  相似文献   

18.
Anandamide (arachidonylethanolamide; AnNH) has important neuromodulatory and immunomodulatory activities. This lipid is rapidly taken up and hydrolyzed to arachidonate and ethanolamine in many organisms. As yet, AnNH inactivation has not been studied in humans. Here, a human brain fatty-acid amide hydrolase (FAAH) has been characterized as a single protein of 67 kDa with a pI of 7.6, showing apparent Km and Vmax values for AnNH of 2.0 +/- 0.2 microM and 800 +/- 75 pmol.min-1.mg of protein-1, respectively. The optimum pH and temperature for AnNH hydrolysis were 9.0 and 37 degreesC, respectively, and the activation energy of the reaction was 43.5 +/- 4.5 kJ.mol-1. Hydro(pero)xides derived from AnNH or its linoleoyl analogues by lipoxygenase action were competitive inhibitors of human brain FAAH, with apparent Ki values in the low micromolar range. One of these compounds, linoleoylethanolamide is the first natural inhibitor (Ki = 9.0 +/- 0.9 microM) of FAAH as yet discovered. An FAAH activity sharing several biochemical properties with the human brain enzyme was demonstrated in human neuroblastoma CHP100 and lymphoma U937 cells. Both cell lines have a high affinity transporter for AnNH, which had apparent Km and Vmax values for AnNH of 0.20 +/- 0.02 microM and 30 +/- 3 pmol.min-1.mg of protein-1 (CHP100 cells) and 0.13 +/- 0.01 microM and 140 +/- 15 pmol.min-1.mg of protein-1 (U937 cells), respectively. The AnNH carrier of both cell lines was activated up to 170% of the control by nitric oxide.  相似文献   

19.
Streptomyces metalloproteinase inhibitor (SMPI) is a small proteinaceous inhibitor which inhibits metalloproteinases such as thermolysin (Ki =1.14 x 10(-10) M). When incubated with the enzyme, it is gradually hydrolyzed at the Cys64-Val65 peptide bond, which was identified as the reactive site by mutational analysis. To achieve a further understanding of the inhibition mechanism, we attempted to resynthesize the cleaved reactive site by using the enzyme catalytic action. The native inhibitor was resynthesized from the modified inhibitor (Ki =2.18 x 10(-8) M) by incubation with a catalytic amount of thermolysin under the same conditions as used for hydrolysis (pH 7.5, 25 degrees C), suggesting that SMPI follows the standard mechanism of inhibition of serine proteinase inhibitors. Temporary inhibition was observed when the native inhibitor and thermolysin were incubated at a 1:100 (mol/mol) enzyme-inhibitor ratio at 37 degrees C. SMPI showed temporary inhibition towards all the enzymes it inhibited. The inhibitory spectrum of SMPI was analyzed with various metalloproteinases based on the Ki values and limited proteolysis patterns. Pseudomonas elastase and Streptomyces griseus metalloproteinase II formed more stable complexes and showed much lower Ki values (approximately 2 pM) than thermolysin. In the limited proteolysis experiments weak inhibitors were degraded by the enzymes. SMPI did not inhibit almelysin, Streptomyces caespitosus neutral proteinase or matrix metalloproteinases. SMPI specifically inhibits metalloproteinases which are sensitive to phosphoramidon.  相似文献   

20.
BACKGROUND: The serine protease prostate-specific antigen (PSA) is a useful clinical marker for prostatic malignancy. PSA is a member of the kallikrein subgroup of the (chymo)trypsin serine protease family, but differs from the prototypical member of this subgroup, tissue kallikrein, in possessing a specificity more similar to that of chymotrypsin than trypsin. We report the use of two strategies, substrate phage display and iterative optimization of natural cleavage sites, to identify labile sequences for PSA cleavage. RESULTS: Iterative optimization and substrate phage display converged on the amino-acid sequence SS(Y/F)Y decreases S(G/S) as preferred subsite occupancy for PSA. These sequences were cleaved by PSA with catalytic efficiencies as high as 2200-3100 M-1 s-1, compared with values of 2-46 M-1 s-1 for peptides containing likely physiological target sequences of PSA from the protein semenogelin. Substrate residues that bind to secondary (non-S1) subsites have a critical role in defining labile substrates and can even cause otherwise disfavored amino acids to bind in the primary specificity (S1) pocket. CONCLUSION: The importance of secondary subsites in defining both the specificity and efficiency of cleavage suggests that substrate recognition by PSA is mediated by an extended binding site. Elucidation of preferred subsite occupancy allowed refinement of the structural model of PSA and should facilitate the development of more sensitive activity-based assays and the design of potent inhibitors.  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