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 15 毫秒
1.
Reversible inhibition of acetylcholinesterase (AChE) from bovine erythrocytes and butyrylcholinesterase (BuChE) from horse blood serum by quaternary diaminoalkyl esters of suberic (D-6), p-phenylenediacetic (PK-139), p-phenylenedipropionic (PK-154 and PK-155), p-phenylenediacrylic (PK-150 and PK-151) and phthalic (PK-105) acids, was studied under the following incubation conditions: pH 7.5, 25 degrees C, 0.1 M KCl. The inhibition kinetics were of a mixed competitive-incompetitive type, the incompetitive component alpha'-having higher values for AChE (0.26-0.60) than for BuChE (0.10-0.20). Diester PK-150 selectively inhibited BuChE (Ki=3.0-10(-6) M); its Ki value for AChE was 4.0-10(-4) M. The other diesters had a stronger inhibitory effect on AChE than on BuChE. High values of alpha' observed during AChE inhibition cannot be interpreted in terms of interaction of those bisquaternary compounds with the anionic site of the acetylated active centre and are probably due to their sorbtion at the peripheral anionic sites. Incompetitive inhibition constants (K'i=Ki/alpha') of BuChE by the diesters PK-139, PK-154 and PK-150 were found to be values of the same order as substrate inhibition constants determined in the course of BuChE hydrolysis of these diesters. Incompetitive inhibition found for the esters studied and substrate inhibition during hydrolysis of these compounds are presumably due to the same mechanism.  相似文献   

2.
Steady-state kinetics for the hydrolysis of benzoylcholine (BzCh) and benzoylthiocholine (BzSCh) by wild-type human butyrylcholinesterase (BuChE) and by the peripheral anionic site mutant D70G were compared. kcat/Km for the hydrolysis of BzSCh was 17-fold and 32-fold lower than that for hydrolysis of BzCh by wild-type and D70G, respectively. The rate-limiting step for hydrolysis of BzCh was deacylation, whereas acylation was rate-limiting for hydrolysis of BzSCh. Wild-type enzyme and the D70G mutant were found to reach steady-state velocity slowly with BzCh as the substrate. At pH 6, the approach to steady-state for both enzymes consisted of a mono-exponential acceleration upon which a set of damped oscillations was superimposed. From pH 7 to 8.5, the approach to steady-state consisted of a simple exponential acceleration. The damped oscillations were analyzed by both a numerical approximation and simulation based on a theoretical model. BuChE-catalyzed hydrolysis of the thiocholine analogue of BzCh showed neither lags nor oscillations, under the same conditions. The frequency and amplitude of the damped oscillations decreased as the BzCh concentration increased. The apparent induction time for the exponential portion of the lag was calculated from the envelope of the damped oscillations or from the smooth lag. Wild-type BuChE showed a hyperbolic increase in induction time as the BzCh concentration increased (tau max = 210 s at pH 6.0). However, the induction time for D70G was constant over the whole range of BzCh concentrations (tau max = 60 s at pH 6.0). Thus, the induction time does not conform to a simple hysteretic model in which there is a slow conformational transition of the enzyme from an inactive form E to an active form E'. No pH-dependence of the induction time was found between pH 6.0 and 8.5 in sodium phosphate buffers of various concentrations (from 1 mm to 1 m). However, increasing the pH tended to abolish the oscillations (increase the damping factor). This effect was more pronounced for D70G than for wild-type. Although the lyotropic properties of phosphate change from chaotropic at pH 6.0 to kosmotropic at pH > 8.0, no effect of phosphate concentration on the oscillations was noticed at the different pH values, suggesting that the oscillations are not related to a pH-dependent Hofmeister effect of phosphate ions. Simulation and theoretical analysis of the oscillatory behaviour of the approach to the steady-state for BuChE led us to propose a model for the hysteresis of BuChE with BzCh. In this model, the substrate-free enzyme is present as an equilibrium mixture of two forms, E and E'. Substrate binds to E and E', but only Epsilon'S makes products. It is proposed that oscillations originate from a time-dependent change in the local concentration, solvation and/or conformation of substrate in the bulk solution. 1H-NMR measurements provided evidence for a slow equilibrium between two BzCh conformers. Binding of the conformationally preferred substrate conformer leads to products.  相似文献   

3.
The rate-limiting step for hydrolysis of the positively charged oxoester benzoylcholine (BzCh) by human butyrylcholinesterase (BuChE) is deacylation (k(3)), whereas it is acylation (k(2)) for hydrolysis of the homologous thioester benzoylthiocholine (BzSCh). Steady-state hydrolysis of BzCh and BzSCh by wild-type BuChE and its peripheral anionic site mutant D70G was investigated at different hydrostatic pressures, which allowed determination of volume changes associated with substrate binding, and the activation volumes for the chemical steps. A differential nonlinear pressure-dependence of the catalytic parameters for hydrolysis of both substrates by both enzymes was shown. Nonlinearity of the plots may be explained in terms of compressibility changes or rate-limiting changes. To distinguish between these two possibilities, enzyme phosphorylation by diisopropylfluorophosphate (DFP) in the presence of substrate (BzSCh) under pressure was studied. There was no pressure dependence of volume changes for DFP binding or for phosphorylation of either wild-type or D70G. Analysis of the pressure dependence for steady-state hydrolysis of substrates, and for phosphorylation by DFP provided evidence that no enzyme compressibility changes occurred during the catalyzed reactions. Thus, the nonlinear pressure dependence of substrate hydrolysis reflects changes in the rate-limiting step with pressure. Change in rate-determining step occurred at a pressure of 100 MPa for hydrolysis of BzCh by wild-type and at 75 MPa for D70G. For hydrolysis of BzSCh the change occurred at higher pressures because k(2) < k(3) at atmospheric pressure for this substrate. Elementary volume change contributions upon initial binding, productive binding, acylation and deacylation were calculated from the pressure differentiation of kinetic constants. This analysis shed light on the molecular events taking place along the hydrolysis pathways of BzCh and BzSCh by wild-type BuChE and the D70G mutant. In addition, volume change differences between wild-type and D70G provided new evidence that residue D70 in the peripheral site controls hydration of the active site gorge and the dynamics of the water molecule network during catalysis. Finally, a steady-state kinetic study of the oxyanion hole mutant (G117H) showed that substitution of the ethereal sulfur for oxygen in the substrate alters the final adjustment of substrate in the active site and stabilization of the acylation transition state.  相似文献   

4.
The role of acetylcholinesterase (AChE) in the termination of the cholinergic response through acetylcholine (ACh) hydrolysis and the involvement of plasma butyrylcholinesterase (BuChE), mainly of hepatic origin, in the metabolism of xenobiotics with ester bonds is well known. Besides, BuChE has a crucial role in ACh hydrolysis, especially when selective anticholinesterases inhibit AChE. Herein, we analyzed the gender-related differences and the circadian changes of rat plasma cholinesterases. Plasma and liver cholinesterase activities were evaluated in control or 2–30-day castrated adult male and female rats. Plasma and liver AChE activities did not differ between genders and were not influenced by sex hormone deprivation. BuChE plasma activity was 7 times greater in female, reflecting gender differences in liver enzyme expression. Castration increased liver and plasma BuChE activity in male, while reduced it in female, abolishing gender differences in enzyme activity. Interestingly, female AChE and BuChE plasma activities varied throughout the day, reaching values 27% and 42% lower, respectively, between 2 p.m. and 6 p.m. when compared to the morning peaks at 8 a.m. Castration attenuated daily female BuChE oscillation. On the other hand, male plasma enzymes remained constant throughout the day. In summary, our results show that liver and plasma BuChE, but not AChE, expression is influenced by sex hormones, leading to high levels of blood BuChE in females. The fluctuation of female plasma BuChE during the day should be taken into account to adjust the bioavailability and the therapeutic effects of cholinesterase inhibitors used in cholinergic-based conditions such Alzheimer's disease.  相似文献   

5.
The aryl-acylamidase (AAA) activity of butyrylcholinesterase (BuChE) has been known for a long time. However, the kinetic mechanism of aryl-acylamide hydrolysis by BuChE has not been investigated. Therefore, the catalytic properties of human BuChE and its peripheral site mutant (D70G) toward neutral and charged aryl-acylamides were determined. Three neutral (o-nitroacetanilide, m-nitroacetanilide, o-nitrophenyltrifluoroacetamide) and one positively charged (3-(acetamido) N,N,N-trimethylanilinium, ATMA) acetanilides were studied. Hydrolysis of ATMA by wild-type and D70G enzymes showed a long transient phase preceding the steady state. The induction phase was characterized by a hysteretic "burst". This reflects the existence of two enzyme states in slow equilibrium with different catalytic properties. Steady-state parameters for hydrolysis of the three acetanilides were compared to catalytic parameters for hydrolysis of esters giving the same acetyl intermediate. Wild-type BuChE showed substrate activation while D70G displayed a Michaelian behavior with ATMA as with positively charged esters. Owing to the low affinity of BuChE for amide substrates, the hydrolysis kinetics of neutral amides was first order. Acylation was the rate-determining step for hydrolysis of aryl-acetylamide substrates. Slow acylation of the enzyme, relative to that by esters may, in part, be due suboptimal fit of the aryl-acylamides in the active center of BuChE. The hypothesis that AAA and esterase active sites of BuChE are non-identical was tested with mutant BuChE. It was found that mutations on the catalytic serine, S198C and S198D, led to complete loss of both activities. The silent variant (FS117) had neither esterase nor AAA activity. Mutation in the peripheral site (D70G) had the same effect on esterase and AAA activities. Echothiophate inhibited both activities identically. It was concluded that the active sites for esterase and AAA activities are identical, i.e. S198. This excludes any other residue present in the gorge for being the catalytic nucleophile pole.  相似文献   

6.
The relationship between structure and the Michaelis-Menten constants (Km) for the papain hydrolysis of a series of 37 N-benzoylglycine esters was investigated. The series studied comprises a wide range of aromatic and aliphatic esters with a 5000-fold variation in their Km constants and essentially constant kcat values. It was found that the variation in the Km constants could be rationalized by the following quantitative structure-activity relationship (QSAR): log 1/Km = 8.13F + 0.33Z + 1.27II3' + 1.95. In this equation F is the field inductive parameter, II3' is the hydrophobic constant for the more lipophilic of the two possible meta substituents and Z is the Van der Waals distance from oxygen through the end of the molecule, in the direction of the 4 position of the aromatic ester moiety.  相似文献   

7.
The purpose of this work was to study the catalytic properties of rat butyrylcholinesterase with benzoylcholine (BzCh) and N-alkyl derivatives of BzCh (BCHn) as substrates. Complex hysteretic behaviour was observed in the approach to steady-state kinetics for each ester. Hysteresis consisted of a long lag phase with damped oscillation. The presence of a long lag phase, with no oscillations, in substrate hydrolysis by rat butyrylcholinesterase was also observed with N-methylindoxyl acetate as substrate. Hysteretic behaviour was explained by the existence of two interconvertible butyrylcholinesterase forms in slow equilibrium, while just one of them is catalytically active. The damped oscillations were explained by the existence of different substrate conformational states and/or aggregates (micelles) in slow equilibrium. Different substrate conformational states were confirmed by 1H-NMR. The K(m) values for substrates decreased as the length of the alkyl chain increased. High affinity of the enzyme for the longest alkyl chain length substrates was explained by multiple hydrophobic interactions of the alkyl chain with amino acid residues lining the active site gorge. Molecular modelling studies supported this interpretation; docking energy decreased as the length of the alkyl chain increased. The long-chain substrates had reduced k(cat) values. Docking studies showed that long-chain substrates were not optimally oriented in the active site for catalysis, thus explaining the slow rate of hydrolysis. The hydrolytic rate of BCH12 and longer alkyl chain esters vs. substrate concentration showed a premature plateau far below V(max). This was due to the loss of substrate availability. The best substrates for rat butyrylcholinesterase were short alkyl homologues, BzCh - BCH4.  相似文献   

8.
K M Rigtrup  D E Ong 《Biochemistry》1992,31(11):2920-2926
Retinol esterified with long-chain fatty acids is a common dietary source of vitamin A. Hydrolysis of these esters in the lumen of the small intestine is required prior to absorption. Bile salt-stimulated retinyl esterase activity was present with purified rat intestinal brush border membrane, with the maximum rate of ester hydrolysis at approximately pH 8, the physiological luminal pH. Taurocholate, a trihydroxy bile salt, stimulated hydrolysis of short-chain fatty acyl retinyl esters more than hydrolysis of long-chain fatty acyl esters. Deoxycholate, a dihydroxy bile salt, primarily stimulated hydrolysis of long-chain esters. Calculated Kms of 0.74 microM for retinyl palmitate (16:0) hydrolysis and 9.6 microM for retinyl caproate (6:0) hydrolysis suggested the presence of two separate activities. Consistent with that, the activity responsible for retinyl caproate hydrolysis could be inactivated to a greater degree than retinyl palmitate hydrolysis by preincubation of the brush border membrane at 37 degrees C for extended times. Brush border membrane from animals who had undergone common duct ligation 48 h prior to tissue collection showed little ability to hydrolyze retinyl caproate but retained 70% of retinyl palmitate hydrolytic activity, compared to sham-operated controls. Thus, two distinguishable retinyl esterase activities were recovered with purified brush border membranes. One apparently originated from the pancreas, was stimulated by trihydroxy bile salts, and preferentially hydrolyzed short-chain retinyl esters, properties similar to cholesterol ester hydrolase, known to bind to the brush border. The other was intrinsic to the brush border, stimulated by both trihydroxy and dihydroxy bile salts, and preferentially hydrolyzed long-chain retinyl esters, providing the majority of activity of the brush border against dietary retinyl esters.  相似文献   

9.
Kinetic constants for the hydrolysis by porcine tissue beta-kallikrein B and by bovine trypsin of a number of peptides related to the sequence of kininogen (also one containing a P2 glycine residue instead of phenylalanine) and of a series of corresponding arginyl peptide esters with various apolar P2 residues have been determined under strictly comparative conditions. kcat and kcat/Km values for the hydrolysis of the Arg-Ser bonds of the peptides by trypsin are conspicuously high. kcat for the best of the peptide substrates, Ac-Phe-Arg-Ser-Val-NH2, even reaches kcat for the corresponding methyl ester, indicating rate-limiting deacylation also in the hydrolysis of a peptide bond by this enzyme. kcat/Km for the hydrolysis of the peptide esters with different nonpolar L-amino acids in P2 is remarkably constant (range 1.7), as it is for the pair of the above pentapeptides with P2 glycine or phenylalanine. kcat for the ester substrates varies fivefold, however, being greatest for the P2 glycine compounds. Obviously, an increased potential of a P2 residue for interactions with the enzyme lowers the rate of deacylation. In contrast to results obtained with chymotrypsin and pancreatic elastase, trypsin is well able to tolerate a P3 proline residue. In the hydrolysis of peptide esters, tissue kallikrein is definitely superior to trypsin. Conversely, peptide bonds are hydrolyzed less efficiently by tissue kallikrein and the acylation reaction is rate-limiting. The influence of the length of peptide substrates is similar in both enzymes and indicates an extension of the substrate recognition site from subsite S3 to at least S'3 of tissue kallikrein and the importance of a hydrogen bond between the P3 carbonyl group and Gly-216 of the enzymes. Tissue kallikrein also tolerates a P3 proline residue well. In sharp contrast to the behaviour of trypsin is the very strong influence of the P2 residue in tissue-kallikrein-catalyzed reactions. kcat/Km varies 75-fold in the series of the dipeptide esters with nonpolar L-amino acid residues in P2, a P2 glycine residue furnishing the worst and phenylalanine the best substrate, whereas this exchange in the pentapeptides changes kcat/Km as much as 730-fold. This behaviour, together with the high value of kcat/Km for Ac-Phe-Arg-OMe of 3.75 X 10(7) M-1 s-1, suggests rate-limiting binding (k1) in the hydrolysis of the best ester substrates.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

10.
Hydrolysis of the neutral substrate N-methylindoxyl acetate (NMIA) by wild-type human butyrylcholinesterase (BuChE) and peripheral site mutants (D70G, Y332A, D70G/Y332A) was found to follow the Michaelis-Menten kinetics. K(m) was 0.14 mM for wild-type, and 0.07-0.16 mM for D70G, Y332A and D70G/Y332A, indicating that the peripheral site is not involved in NMIA binding. The values of k(cat) were of the same order for all enzymes: 12,000-18,000 min(-1). Volume changes upon substrate binding (-DeltaV(K(m))) and the activation volumes (DeltaV++(k(cat)) associated with hydrolysis of NMIA were calculated from the pressure dependence of the catalytic constants. Values of -DeltaV(K(m)) indicate that NMIA binds to an aromatic residue, presumed to be W82, the active site binding locus. Binding is accompanied by a release of water molecules from the gorge. Residue 70 controls the number of water molecules that are released upon substrate binding. The values of DeltaV++(k(cat)), which are positive for wild-type and faintly positive for D70G, clearly indicate that the catalytic steps are accompanied by re-entry of water into the gorge. Results support the premise that residue D70 is involved in the conformational stabilization of the active site gorge and in control of its hydration. A slow transient, preceding the steady state, was seen on a time scale of several minutes. The induction time rapidly increased with NMIA concentration to reach a limit at substrate saturation. Much shorter induction times (<1 min) were seen for hydrolysis of benzoylcholine (BzCh) by wild-type BuChE and for hydrolysis of butyrylthiocholine (BuSCh) by the active site mutants E197Q and E197Q/G117H. This slow transient was interpreted in terms of hysteresis without kinetic cooperativity. The hysteretic behavior of BuChE results from a slow conformational equilibrium between two enzyme states E and E'. NMIA binds only to the primed form E'. Kosmotropic salts and hydrostatic pressure were found to shift the equilibrium toward E'. The E-->E' transition is accompanied by a negative activation volume (DeltaV++(0)= -45+/-10 ml/mol), and the E' form is more compact than E. Hydration water in the gorge of E' appears to be more structured than in the unprimed form.  相似文献   

11.
Porcine pancreatic lipase (PPL)-catalyzed enantioselective hydrolysis of N-benzyloxycarbonyl-dl-amino acid esters (Z-dl-AA-ORs) was studied for the optical resolution of a variety of non-protein amino acids. The ester moiety (R) of the substrate affected the rate of hydrolysis significantly. The glyceryl (Gl) and carbamoylmethyl (Cam) esters were found to be highly reactive substrates. The hydrolysis of the Gl esters (Z-dl-AA-OGls) of both aliphatic and aromatic amino acids was examined in acetonitrile containing 70% (v/v) of 0.02 M phosphate buffer (pH 7.0) at 30°C. With all amino acids tested, the corresponding l-enantiomers were hydrolyzed preferentially. PPL favored aromatic amino acids, such as phenylalanine and p-chlorophenylalanine, leading to completion of the hydrolysis within 20 min with excellent enantioselectivities (E>100). The PPL-catalyzed hydrolysis of the corresponding Cam esters (Z-dl-AA-OCams) was also examined under the same reaction conditions. Although the hydrolysis of the Cam esters was rapid, the l-enantioselectivities were rather poor with aromatic amino acids, such as 2-phenylglycine and homophenylalanine.  相似文献   

12.
S Naderi  K Fotherby 《Steroids》1983,41(3):397-417
The hydrolysis of 108 esters of norethisterone (17 alpha-ethynyl-17 beta-hydroxyestr-4-en-3-one) was studied in vitro using a rabbit liver preparation. Introduction of a double or triple bond into a straight-chain ester did not inhibit hydrolysis but a marked reduction in hydrolysis was produced on replacement of a methylene group by an oxygen atom. Hydrolysis was inhibited by substituents at C2 of the ester chain except in short chain esters. Cyclopropylcarboxylate and cyclobutylcarboxylate were readily hydrolysed and introduction of a furan ring into the side-chain did not affect hydrolysis. No hydrolysis occurred with a cholesteryl carbonate ester or with a pentamethyldisilyloxy ether. Forty-nine esters of levonorgestrel (13 beta-ethyl-17 alpha-ethynyl-17 beta-hydroxygon-4-en-3-one) were also studied. In general, the pattern of hydrolysis for these esters was similar to that of the norethisterone esters. However, with few exceptions the levonorgestrel esters were hydrolyzed more slowly. For those esters for which information regarding the biological activity was available, there was no correlation between the potency of the esters and their rate of hydrolysis in vitro.  相似文献   

13.
Leaf waxes from spring wheat varieties Selkirk and Manitou contain hydrocarbons (6%, 10%), long chain esters (14%, 13%), free acids (5%, 8%), free alcohols (19%, 21%), β-diketone (16%, 20%), hydroxy β-diketones (8%, 10%), unidentified gum (29%, 16.5%) and minor amounts of diol diesters, glycerides and aldehydes. The major hydrocarbon is nonacosane and major esters are octacosyl esters of C14–C32 acids but C20 and C22 alcohol esters of trans 2-docosenoic and tetracosenoic acids are also present (Selkirk 20%, Manitou 10% of total esters). Previously unknown trans 2-docosen-1-ol is present as an ester (Selkirk 5%, Manitou 2.5% of total esters). Free acids are C14–C32 acids and trans 2-docosenoic and tetracosenoic acids (Selkirk 30%, Manitou 9% of free acids). Octacosanol is the principal free alcohol. Hentriacontane-14,16-dione is the β-diketone and the hydroxy β-diketones are a 1:1 mixture of 8- and 9- hydroxyhentriacontane-14,16-diones.  相似文献   

14.
Trimethyl and triethyl esters of tricarballylic acid and citric acid were hydrolysed with porcine liver esterase(PLE) to the isomeric diesters. In all cases the hydrolysis took place with poor regioselectivity (maximum 50% excess). However, the hydrolysis of trimethyl and triethyl esters of tricarballylic acid and of the triethyl ester of citric acid with subtilisin was absolutely regioselective and the symmetric 1,5-diester was obtained.  相似文献   

15.
(Hetero)aromatic mono- and diesters of tropine and nortropine were prepared. Modulation of [3H]strychnine binding to glycine receptors of rat spinal cord was examined with a ternary allosteric model. The esters displaced [3H]strychnine binding with nano- or micromolar potencies and strong negative cooperativity. Coplanarity and distance of the ester moieties of diesters affected the binding affinity being nanomolar for isophthaloyl-bistropane and nortropeines. Nortropisetron had the highest affinity (K(A) approximately 10 nM). Two esters displayed negative cooperativity with glycine in displacement, while three esters of low-affinity and nortropisetron exerted positive cooperativity with glycine.  相似文献   

16.
Residues and coal fractions that remained after the biosolubilization of Rhenish brown coal by strains of Lentinula edodes and Trametes versicolor have been studied by Curie-point pyrolysis/gas chromatography/mass spectrometry using tetraethylammonium hydroxide (NEt4OH) at 610 °C. To differentiate methyl derivatives of esters and ethers from free or bound hydroxyl and carboxyl groups NEt4OH was used in the thermochemolysis experiments instead the commonly used tetramethylammonium hydroxide. A comparison of humic acid fractions before and after fungal attack shows considerable alteration of the soluble macromolecules of coal. Depending on the coal fraction studied and the fungi used, the assortment of fatty acid esters released during the pyrolysis varies significantly. Furthermore, dicarbonic acid ethyl diesters as well as ethyl derivatives of aromatic ethers and acids yield information about humic acid structure and the biosolubilization of brown coal. Variations in the mixture produced are possibly caused by differences in the pattern of extracellular enzymes secreted that attack the macromolecular structural elements of brown coal. Therefore pyrolysis of native and microbiologically altered geomacromolecules using NEt4OH allows one to differentiate between free hydroxyl groups as well as substances that are attached to humic substances via ester or ether bridges, and their methylated counterparts. Received: 13 July 1998 / Received revision: 12 October 1998 / Accepted: 16 October 1998  相似文献   

17.
A new method was developed for analyzing the steric configuration of hydroxyeicosatetraenoates (HETEs) and other hydroxy fatty acids. Racemic HETE methyl esters were reacted with either benzoyl or naphthoyl chloride in pyridine and the resulting aromatic ester derivatives purified by reversed phase HPLC and subsequently chromatographed on a chiral stationary phase HPLC column [(R)-(-)-N-3,5-dinitrobenzoyl-alpha-phenylglycine)]. In contrast to the enantiomers of the underivatized HETE methyl esters which were only partially resolved, the enantiomers of their aromatic ester derivatives were completely separable on this chiral phase. Chiral HETEs can be retrieved from the aromatic derivatives by alkaline hydrolysis. Thus, this method has both analytical and preparative applications.  相似文献   

18.
The kinetics of hydrolysis of dipropanoylmorphine (DPM) and dibutanoylmorphine (DBM) in human blood fractions and for diacetylmorphine (DAM) and DBM in rat blood fractions were investigated. In each case the hydrolysis of morphine diesters terminated with the production of the corresponding 6-monoester derivative. Generally, decreases in Km and Vmax were observed for the plasma, red blood cell (RBC) cytosol, and RBC membrane esterases responsible for morphine diester hydrolysis as the alkyl chain length of the ester moiety increased. This resulted in an overall decrease in the rate of hydrolysis of morphine diesters by human or rat blood with longer chain homologs of DAM. The analgesic potency and duration of morphine, DAM, and DBM were assessed at various i.v. dosages in the rat by means of the tail-flick latency test. A comparison of equianalgesic doses of morphine, DAM, and DBM indicated that DAM and DBM were 11.5 and 6 times as potent and 0.8 and 1.2 times as long acting, respectively, as morphine.  相似文献   

19.
A serine proteinase (ycaB) from the yeast Candida albicans A.T.C.C. 10261 was purified to near homogeneity. The enzyme was almost indistinguishable from yeast proteinase B (EC 3.4.21.48), and an Mr of 30,000 for the proteinase was determined by SDS/polyacrylamide-gel electrophoresis. The initial site of hydrolysis of the oxidized B-chain of insulin, by the purified proteinase, was the Leu-Tyr peptide bond. The preferential degradation at this site, analysed further with N-blocked amino acid ester and amide substrates, demonstrated that the specificity of the proteinase is determined by an extended substrate-binding site, consisting of at least three subsites (S1, S2 and S'1). The best p-nitrophenyl ester substrates were benzyloxycarbonyl-Tyr p-nitrophenyl ester (kcat./Km 3,536,000 M-1 X S-1), benzyloxycarbonyl-Leu p-nitrophenyl ester (kcat./Km 2,250,000 M-1 X S-1) and benzyloxycarbonyl-Phe p-nitrophenyl ester (kcat./Km 1,000,000 M-1 X S-1) consistent with a preference for aliphatic or aromatic amino acids at subsite S1. The specificity for benzyloxycarbonyl-Tyr p-nitrophenyl ester probably reflects the binding of the p-nitrophenyl group in subsite S'1. The presence of S2 was demonstrated by comparison of the proteolytic coefficients (kcat./Km) for benzyloxycarbonyl-Ala p-nitrophenyl ester (825,000 M-1 X S-1) and t-butyloxycarbonyl-Ala p-nitrophenyl ester (333,000 M-1 X S-1). Cell-free extracts contain a heat-stable inhibitor of the proteinase.  相似文献   

20.
Many toxic secondary metabolites used for defense are also toxic to the producing organism. One important way to circumvent toxicity is to store the toxin as an inactive precursor. Several sulfated diesters of the diarrhetic shellfish poisoning (DSP) toxin okadaic acid have been reported from cultures of various dinoflagellate species belonging to the genus Prorocentrum. It has been proposed that these sulfated diesters are a means of toxin storage within the dinoflagellate cell, and that a putative enzyme mediated two-step hydrolysis of sulfated diesters such as DTX-4 and DTX-5 initially leads to the formation of diol esters and ultimately to the release of free okadaic acid. However, only one diol ester and no sulfated diesters of DTX-1, a closely related DSP toxin, have been isolated leading some to speculate that this toxin is not stored as a sulfated diester and is processed by some other means. DSP components in organic extracts of two large scale Prorocentrum lima laboratory cultures have been investigated. In addition to the usual suite of okadaic acid esters, as well as the free acids okadaic acid and DTX-1, a group of corresponding diol- and sulfated diesters of both okadaic acid and DTX-1 have now been isolated and structurally characterized, confirming that both okadaic acid and DTX-1 are initially formed in the dinoflagellate cell as the non-toxic sulfated 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