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.
1. To determine kinetic parameters (Vmax, K(m)) for cytochrome P450 (CYP) mediated metabolic pathways, nonlinear least squares regression is commonly used to fit a model equation (e.g., Michaelis Menten [MM]) to sets of data points (reaction velocity vs substrate concentration). This method can also be utilized to determine the parameters for more complex mechanisms involving allosteric or multi-enzyme systems. Akaike's Information Criterion (AIC), or an estimation of improvement of fit as successive parameters are introduced in the model (F-test), can be used to determine whether application of more complex models is helpful. To evaluate these approaches, we have examined the complex enzyme kinetics of amitriptyline (AMI) N-demethylation in vitro by human liver microsomes. 2. For a 15-point nortriptyline (NT) formation rate vs substrate (AMI) concentration curve, a two enzyme model, consisting of one enzyme with MM kinetics (Vmax = 1.2 nmol min-1 mg-1, K(m) = 24 microM) together with a sigmoidal component (described by an equation equivalent to the Hill equation for cooperative substrate binding; Vmax = 2.1 nmol min-1 mg-1, K' = 70 microM; Hill exponent n = 2.34), was favoured according to AIC and the F-test. 3. Data generated by incubating AMI under the same conditions but in the presence of 10 microM ketoconazole (KET), a CYP3A3/4 inhibitor, were consistent with a single enzyme model with substrate inhibition (Vmax = 0.74 nmol min-1 mg-1, K(m) = 186 microM, K1 = 0.0028 microM-1). 4. Sulphaphenazole (SPA), a CYP2C9 inhibitor, decreased the rate of NT formation in a concentration dependent manner, whereas a polyclonal rat liver CYP2C11 antibody, inhibitory for S-mephenytoin 4'-hydroxylation in humans, had no important effect on this reaction. 5. Incubation of AMI with 50 microM SPA resulted in a curve consistent with a two enzyme model, one with MM kinetics (Vmax = 0.72 nmol min-1 mg-1, K(m) = 54 microM) the other with 'Hill-kinetics' (Vmax = 2.1 nmol min-1 mg-1, K' = 195 microM; n = 2.38). 6. A fourth data-set was generated by incubating AMI with 10 microM KET and 50 microM SPA. The proposed model of best fit describes two activities, one obeying MM-kinetics (Vmax = 0.048 nmol min-1 mg-1, K(m) = 7 microM) and the other obeying MM kinetics but with substrate inhibition (Vmax = 0.8 nmol min-1 mg-1, K(m) = 443 microM, K1 = 0.0041 microM-1). 7. The combination of kinetic modelling tools and biological data has permitted the discrimination of at least three CYP enzymes involved in AMI N-demethylation. Two are identified as CYP3A3/4 and CYP2C9, although further work in several more livers is required to confirm the participation of the latter.  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.
BACKGROUND: Canine ventricles have been reported to contain a cathepsin D-like kininogenase, which might confer protection on the heart during ischaemia. The aim of this study was to investigate the presence and levels of a similar kininogenase in normal and ischaemic rat hearts. METHODS: Aqueous extracts of rat ventricles were tested for the ability to release bradykinin-like immunoreactivity from human low-molecular-weight kininogen and high-molecular-weight kininogen at acidic pHs. The enzymes involved were separated using gel filtration followed by the testing of fractions for cleavage of D-Val-Leu-Arg-pNA and low-molecular-weight kininogen. Extracts from normal and ischaemic ventricles were compared for the ability to release bradykinin-like immunoreactivity from low-molecular-weight kininogen. Kinin levels in mixed venous blood were compared before and after ischaemia. By assessing their effect on isolated oestrous rat uteri and on protease inhibitors, further characterization of acidic kininogenases in the extracts was performed. RESULTS: Extracts of rat ventricles released bradykinin-like immunoreactivity only from low-molecular-weight kininogen. Using the isolated oestrous rat uterus, gel filtration and protease inhibitors, the enzyme involved was identified as a cathepsin D-like enzyme with an optimum pH of 4.7 and a molecular weight of 42.8 +/- 4.9 kDa. It is an arginine amidase and releases bradykinin-like immunoreactivity from low-molecular-weight kininogen. Ischaemia reduced the amount of bradykinin-like immunoreactivity released by the ventricular extract (P < or = 0.05) and increased levels of free kinin in venous blood from the right atrium. CONCLUSION: Rat ventricles contain a cathepsin D-like acidic protease that cleaves low-molecular-weight kininogen to release bradykinin-like immunoreactivity. The acidic protease may protect the heart during ischaemia.  相似文献   

4.
Carbonic anhydrase VII (CA VII) appears to be the most highly conserved of the active mammalian carbonic anhydrases. We have characterized the catalytic activity and inhibition properties of a recombinant murine CA VII. CA VII has steady-state constants similar to two of the most active isozymes of carbonic anhydrase, CA II and IV; also, it is very strongly inhibited by the sulfonamides ethoxzolamide and acetazolamide, yielding the lowest Ki values measured by the exchange of 18O between CO2 and water for any of the mammalian isozymes of carbonic anhydrase. The catalytic measurements of the hydration of CO2 and the dehydration of HCO3- were made by stopped-flow spectrophotometry and the exchange of 18O using mass spectrometry. Unlike the other isozymes of this class of CA, for which Kcat/K(m) is described by the single ionization of zinc-bound water, CA VII exhibits a pH profile for Kcat/K(m) for CO2 hydration described by two ionizations at pKa 6.2 and 7.5, with a maximum approaching 8 x 10(7) M-1 s-1. The pH dependence of kcat/K(m) for the hydrolysis of 4-nitrophenyl acetate could also be described by these two ionizations, yielding a maximum of 71 M-1 s-1 at pH > 9. Using a novel method that compares rates of 18O exchange and dehydration of HCO3-, we assigned values for the apparent pKa at 6.2 to the zinc-bound water and the pKa of 7.5 to His 64. The magnitude of Kcat, its pH profile, 18O-exchange data for both wild-type and a H64A mutant, and inhibition by CuSO4 and acrolein suggest that the histidine at position 64 is functioning as a proton-transfer group and is responsible for one of the observed ionizations. A truncation mutant of CA VII, in which 23 residues from the amino-terminal end were deleted, has its rate constant for intramolecular proton transfer decreased by an order of magnitude with no change in Kcat/K(m). This suggests a role for the amino-terminal end in enhancing proton transfer in catalysis by carbonic anhydrase.  相似文献   

5.
A proteinase accumulated in breast secretions from women with breast cancer has been characterised. Inhibition of the proteolytic activity of breast secretions by pepstatin A showed that the main enzyme involved was an aspartyl proteinase. Determination of its cleavage specificity by SDS-PAGE and amino acid sequence analysis revealed that it was identical to that of cathepsin D, an aspartyl proteinase suggested to be involved in breast cancer development. The identity between both proteins was further confirmed by immunological analysis with monoclonal antibodies against cathepsin D. Quantification of cathepsin D in nipple fluids from 41 women with benign or malignant breast diseases and from 19 control women without breast pathology revealed the presence of variable amounts of this proteinase. The average concentration of cathepsin D in breast secretions from cancer-bearing breasts was 7.2 +/- 2.2 fmol micrograms of protein, which was significantly higher than those of nipple fluids from control women (2.9 +/- 0.6 fmol micrograms-1) (P = 0.04) or from patients with benign breast diseases (2.1 +/- 0.3 fmol micrograms-1) (P = 0.004). Though the number of cancer patients studied was small (n = 21), no correlations were found with cytosolic concentrations of cathepsin D or oestrogen receptors, neither with other parameters such as tumour size, histological grade, axillary node involvement or menopausal status.  相似文献   

6.
Two intramolecularly quenched fluorogenic peptides containing o-aminobenzoyl (Abz) and ethylenediamine 2,4-dinitrophenyl (EDDnp) groups at amino- and carboxyl-terminal amino acid residues, Abz-DArg-Arg-Leu-EDDnp (Abz-DRRL-EDDnp) and Abz-DArg-Arg-Phe-EDDnp (Abz-DRRF-EDDnp), were selectively hydrolyzed by neutral endopeptidase (NEP, enkephalinase, neprilysin, EC 3.4.24.11) at the Arg-Leu and Arg-Phe bonds, respectively. The kinetic parameters for the NEP-catalyzed hydrolysis of Abz-DRRL-EDDnp and Abz-DRRF-EDDnp were K(m) = 2.8 microM, kcat = 5.3 min-1, kcat/K(m) = 2 min-1 microM-1 and K(m) = 5.0 microM, kcat = 7.0 min-1, kcat/K(m) = 1.4 min-1 microM-1, respectively. The high specificity of these substrates was demonstrated by their resistance to hydrolysis by metalloproteases [thermolysin (EC 3.4.24.2), angiotensin-converting enzyme (ACE; EC 3.4.24.15)], serineproteases [trypsin (EC 3.4.21.4), alpha-chymotrypsin (EC 3.4.21.1)] and proteases present in tissue homogenates from kidney, lung, brain and testis. The blocked amino- and carboxyl-terminal amino acids protected these substrates against the action of aminopeptidases, carboxypeptidases and ACE. Furthermore, DR amino acids ensured total protection of Abz-DRRL-EDDnp and Abz-DRRF-EDDnp against the action of thermolysin and trypsin. Leu-EDDnp and Phe-EDDnp were resistant to hydrolysis by alpha-chymotrypsin. The high specificity of these substrates suggests their use for specific NEP assays in crude enzyme preparations.  相似文献   

7.
1. The present experiments were undertaken in order to characterize further the apparently irreversible inhibition of the contraction of depolarized rat aorta caused by lacidipine, a 1,4-dihydropyridine calcium antagonist. 2. We studied the effect of lacidipine on contraction evoked by 100 mM KCl solution in rat aorta, treated by N omega-nitro-L-arginine (0.1 mM), an inhibitor of nitric oxide (NO) synthesis. We compared the effect of prolonged depolarization on lacidipine and (+)-isradipine inhibition and the reversal of this inhibition after washout in the absence of dihydropyridines. Assuming that the onset of lacidipine-evoked inhibition was a pseudo-first order association kinetics, we estimated the dissociation rate constant (k-1 = 0.031 min-1), the association rate constant (k1 = 2.70 x 10(8) M-1 min-1) and the dissociation constant (KD = k-1/k1 = 115 pM) which was close to the IC50 value in steady-state conditions (160 pM). 3. The inhibitory effects of lacidipine and (+)-isradipine on rat aorta contraction were reversibly enhanced after preincubation with the drug in a 40 mM KCl-solution. Washout with drug-free 40 mM K(+)-depolarizing solution reversed inhibition in the (+)-isradipine-treated preparations, but not in the lacidipine-treated ones. 4. Radioligand binding studies were performed with [3H]-lacidipine and [3H]-isradipine in microsomes from rat aorta and rat ileum. Both ligands bound to a homogeneous population of binding sites (for[3H]-lacidipine: KD = 23 +/- 2.6 pM, Bmax = 380 +/- 21 fmol mg-1 protein in membranes from aorta; KD =23 +/- 3.1 pM, Bmax = 790 +/- 60 fmol mg-1 protein in membranes from ileum; for [3H]-isradipine:KD = 140 +/- 46 pM, Bmax = 350 +/- 64 fmol mg-1 protein in membrane from aorta; KD = 68 +/- 14 pM,Bmax = 760 +/- 75 fmol mg-1 protein in membranes from ileum). After isotopic dilution, [3H]-lacidipine and [3H]-isradipine dissociated according to a monoexponential kinetics. In membranes from ileum, the calculated dissociation rate constants (kappa_ 1) were 0.0257 min-1 and 0.0595 min-1, for [3H]-lacidipine and[3H]-isradipine, respectively.5. The non specific binding of [3H]-lacidipine and [3H]-isradipine, was measured in intact rat aorta preparations incubated under the conditions of the functional experiments, in the presence of nifedipine(1 microM). After incubation with [3H]-lacidipine 77.6 +/- 1.9 pM for 2 h the concentration of drug in the tissue was 15.15 +/- 1.18 fmol mg-1 w.wt. and still amounted to 7.24 +/- 0.61 fmol mg-1 w.wt. after 3.5 h washout in drug-free solution. After incubation with [3H]-isradipine 47.2 +/- 0.4 pM for 2 h it was 2.26 +/-0.07 fmol mg-1 w.wt. and was undetectable after 3.5 h washout in a drug-free solution.6. It is concluded that lacidipine interacts reversibly with dihydropyridine binding sites and that the apparent irreversible inhibition of contraction in depolarized preparations could be related to a nonspecific binding in a tissue compartment different from the plasma membrane.  相似文献   

8.
We synthesized a novel potent alpha-chymotrypsin inactivator, 2,2-dimethyl-3-(N-4-cyanobenzoyl) amino-5-phenyl pentanoic anhydride, which fulfilled the criteria of a mechanism-based inactivator: first-order kinetics, irreversibility, saturation kinetics and substrate protection. The inactivation rate constant (kinact) and the enzyme-inhibitor dissociation constant (KI) were calculated to be 0.017s-1 and 0.071 microM, respectively (kinact/KI = 242,000 M-1s-1). These kinetic parameters indicate that this compound is one of the most powerful alpha-chymotrypsin inactivators ever reported. The average number of alpha-chymotrypsin turnovers per inactivation (partition ratio) was calculated to be 1, which indicates that it is a stoichiometrically ideal inactivator of alpha-chymotrypsin. We compared the IC50 values of this compound with those of several chymotrypsin-like serine proteinases (bovine alpha-chymotrypsin, recombinant human chymase and human neutrophil cathepsin G) and a metallo proteinase, rabbit angiotensin converting enzyme (ACE). Our compound, 2,2-dimethyl-3-(N-4-cyanobenzoyl) amino-5-phenyl pentanoic anhydride, inhibited bovine alpha-chymotrypsin potently (IC50 = 1.0 (+/- 0.2) x 10(-9) M) as well as other chymotrypsin-like serine proteinase; recombinant human chymase (IC50 = 7.0 (+/- 1.0) x 10(-8) M) and human neutrophil cathepsin G (IC50 = 1.8 (+/- 0.2) x 10(-7) M). However, rabbit ACE was not inhibited by this compound (IC50 > 1 x 10(-4) M).  相似文献   

9.
During anaerobic growth of Klebsiella pneumoniae on citrate, 9.4 mmol of H2/mol of citrate (4-kPa partial pressure) was formed at the end of growth besides acetate, formate, and CO2. Upon addition of NiCl2 (36 microM) to the growth medium, hydrogen formation increased about 36% to 14.8 mmol/mol of citrate (6 kPa), and the cell yield increased about 15%. Cells that had been harvested and washed under anoxic conditions exhibited an H2-dependent formation of NAD(P)H in vivo. The reduction of internal NAD(P)+ was also achieved by the addition of formate. In crude extracts, the H2:NAD+ oxidoreductase activity was 0.13 micromol min-1 mg-1, and 76% of this activity was found in the washed membrane fraction. The highest specific activities of the membrane fraction were observed in 50 mM potassium phosphate, with 1.6 micromol of NADPH formed min-1 mg-1 at pH 7.0 and 1.7 micromol of NADH formed min-1 mg-1 at pH 9.5. In the presence of the protonophore carbonyl cyanide m-chlorophenylhydrazone and the Na+/H+ antiporter monensin, the H2-dependent reduction of NAD+ by membrane vesicles decreased only slightly (about 16%). The NADP+- or NAD+-reducing hydrogenases were solubilized from the membranes with the detergent lauryldimethylamine-N-oxide or Triton X-100. NAD(P)H formation with H2 as electron donor, therefore, does not depend on an energized state of the membrane. It is proposed that hydrogen which is formed by K. pneumoniae during citrate fermentation is recaptured by a novel membrane-bound, oxygen-sensitive H2:NAD(P)+ oxidoreductase that provides reducing equivalents for the synthesis of cell material.  相似文献   

10.
The capsid of canine parvovirus (CPV) was assayed for susceptibility to proteases and for structural variation. The natural cleavage of VP2 to VP3 in CPV full (DNA containing) particles recovered from tissue culture occurred within the sequence Arg-Asn-Glu-Arg Ala-Thr. Trypsin, chymotrypsin, bromelain, and cathepsin B all cleaved >90% of the VP2 to VP3 in full but not in empty capsids and did not digest the capsid further. Digestion with proteinase K, Pronase, papain, or subtilisin cleaved the VP2 to VP3 and also cleaved at additional internal sites, causing particle disintegration and protein degradation. Several partial digestion products produced by proteinase K or subtilisin were approximately 31-32.5 kDa, indicating cleavage within loop 3 of the capsid protein as well as other sites. Protease treatment of capsids at pH 5.5 or 7.5 did not significantly alter their susceptibility to digestion. The isoelectric point of CPV empty capsids was pH 5.3, and full capsids were 0.3 pH more acidic, but after proteolysis of VP2 to VP3, the pI of the full capsids became the same as that of the empty capsids. Antibodies against various capsid protein sequences showed the amino termini of most VP2 molecules were on the outside of full but not empty particles, that the VP1-unique sequence was internal, and that the capsid could be disintegrated by heat or urea treatment to expose the internal sequences. Capsids added to cells were localized within the cell cytoplasm in vesicles that appeared to be lysosomes. Microinjected capsids remained primarily in the cytoplasm, although a small proportion was observed to be in the nucleus after 2 h. After CPV capsids labeled with [35S]methionine were bound to cells at 0 degrees C and the cells warmed, little cleavage of VP1 or VP2 was observed even after prolonged incubation. Inoculation of cells with virus in the presence of proteinase inhibitors did not significantly reduce the infection.  相似文献   

11.
Cathepsin D, a lysosomal aspartic protease, has been purified from porcine liver using a combination of pepstatin-A agarose and Affi-Gel Blue affinity chromatography, followed by size-exclusion chromatography. The purified protein consists of two polypeptide chains of 15 and 30 kDa, and has an isoelectric point of 6.8. Porcine liver cathepsin D has maximum activity at pH 2.5-3.0 as determined by its activity against hemoglobin, with a Kcat of 14.3 s-1 and a kcat/KM of 2.70 x 10(6) s-1M-1 as determined by the hydrolysis of a fluorogenic peptide substrate.  相似文献   

12.
The reversed-phase HPLC separation of fluorescent o-phthalaldehyde (OPA) derivatives has been applied to the assay of hepatic gamma-glutamylcysteine and glutathione (GSH) levels and the enzymes producing these peptides. The method has been compared to the assay using monobromobimane (MB) as the derivatizing agent. The OPA method has the advantage of faster derivatization, the lack of need to adjust the pH, isocratic separation and selectivity for GSH and gamma-glutamylcysteine. The MB method requires pH adjustment following derivatization and gradient elution chromatography. MB is also non-selective, yielding fluorescent derivatives of all biological thiols and more interfering peaks on the chromatogram. MB-based analyses are also approximately sixty times more expensive per sample. MB yields fluorescent degradation products on exposure to light. OPA adducts are stable for up to ten days when stored at -20 degrees C. OPA detection is sensitive to 12.5 pmol in the sample, at a signal-to-noise ratio of 2.5. The two methods correlate well. Hepatic gamma-glutamylcysteine synthetase in the same liver preparation was found to be 4.85 +/- 0.47 nmol min-1 mg-1 protein by the OPA method and 4.42 +/- 0.52 nmol min-1 mg-1 protein by the MB method. GSH concentrations were found to be 90.4 +/- 6.5 nmol/mg protein for the OPA method and 92.5 +/- 3.4 for the MB method.  相似文献   

13.
A new cysteine proteinase, salmon miltpain, was isolated and purified from the milt of chum salmon (Oncorhynchus keta). Native molecular mass was estimated as 67,000 by gel filtration column chromatography (Shodex WS2003) and 22,300 by SDS-polyacrylamide gel electrophoresis. Isoelectoric point was determined to be 3.9 by isoelectric focusing. The first 15 amino acid residues in the N-terminal region were LPSFLY-AEMVGYNIL. The cysteine proteinase, which had a pH optimum of 6.0 for Z-Arg-Arg-MCA hydrolysis, required a thiol-reducing reagent for activation and was inhibited by E-64, iodacetamide, CA-074 Me, TLCK, TPCK and ZPCK. The cysteine proteinase exhibited unique substrate specificity toward paired basic residues such as Lys-Arg, Arg-Arg at the subsites of P2-P1 and had a K(m) of 16.3 microM and kcat of 20.3 s-1 with Z-Arg-Arg-MCA as substrate and a K(m) of 52.9 microM and kcat of 1.79 s-1 with Z-Phe-Arg-MCA. This proteinase was found to considerably hydrolyze basic proteins such as histone, salmine and clupaine but not milk casein.  相似文献   

14.
Fe(III) uptake by the iron-delivery and iron-scavenging protein, hen ovotransferrin has been investigated in vitro between pH 6.5 and 9. In the absence of any ferric chelate, apo-ovotransferrin loses two protons with K1a = 50 +/- 1 nM and K2a = 4.0 +/- 0.1 nM. These acid-base equilibria are independent of the interaction of the protein with bicarbonate. The interaction with bicarbonate occurs with two different affinity constants, KC = 9.95 +/- 0.15 mM and KN = 110 +/- 10 mM. FeNAc3 exchanges its Fe(III) with the C-site of the protein in interaction with bicarbonate, direct rate constants k1 = 650 +/- 25 M-1 s-1, reverse rate constant k-1 = (6.0 +/- 0.1) x 10(3) M-1 s-1 and equilibrium constant K1 = 0.11 +/- 0.01. This iron-protein intermediate loses then a single proton, K3a = 3.50 +/- 0.35 nM, and undergoes a first change in conformation followed by a two or three proton loss, first order rate constant k2 = 0.30 +/- 0.01 s-1. This induces a new modification in conformation followed by the loss of one or two protons, first order rate constant k3 = (1.50 +/- 0.05) x 10(-2) s-1. These modifications in the monoferric protein conformation are essential for iron uptake by the N-site of the protein. In the last step, the monoferric and diferric proteins attain their final state of equilibrium in about 15,000 s. The overall mechanism of iron uptake by ovotransferrin is similar but not identical to those of serum transferrin and lactoferrin. The rates involved are, however, closer to lactoferrin than serum transferrin, whereas the affinities for Fe(III) are lower than those of serum transferrin and lactoferrin. Does this imply that the metabolic function transferrins is more related to kinetics than to thermodynamics?  相似文献   

15.
Ester and amide derivatives of alpha-azaglycine (carbazic acid, H2NNHCOOH), alpha-azaalanine, and alpha-azaphenylalanine (i.e., Ac-l-Phe-NHN(R)CO-X, where X = H, CH3, or CH2Ph, respectively) were synthesized and evaluated as inhibitors of the cysteine proteinases papain and cathepsin B. The ester derivatives inactivated papain and cathepsin B at rates which increased dramatically with leaving group hydrophobicity and electronegativity. For example, with 8 (R = H, X = OPh) the apparent second-order rate constant for papain inactivation was 67 600 M-1 s-1. Amide and P1-thioamide derivatives do not inactivate papain, nor are they substrates; instead they are weak competitive inhibitors (0.2 mM < Ki < 4 mM). Inactivation of papain involves carbamoylation of the enzyme, as demonstrated by electrospray mass spectrometry. Active site titration indicated a 1:1 stoichiometry for the inactivation of papain with 8, and both inactivated papain and cathepsin B are highly resistant to reactivation by dialysis (t1/2 > 24 h at 4 degrees C). Azaalanine derivatives Ac-L-Phe-NHN(CH3)CO-X inactivate papain ca. 400- 900-fold more slowly than their azaglycine analogues, consistent with the planar configuration at Nalpha of the P1 residue and the very substantial stereoselectivity of papain for L- vs D- residues at the P1 position of its substrates. Azaglycine derivative 9 (R = H, X = OC6H4NO2-p) inactivates papain extremely rapidly (>70 000 M-1 s-1), but it also decomposes rapidly in buffer with release of nitrophenol (kobs = 0.13 min-1); under the same conditions 8 shows <7% hydrolysis over 24 h. This nitrophenol release probably involves cyclization to an oxadiazolone since 17 (R = CH3, X = OC6H4NO2-p), which cannot form an isocyanate, releases nitrophenol almost as rapidly (kobs = 0.028 min-1). Cathepsin C, another cysteine proteinase with a rather different substrate specificity (i.e., aminopeptidase), was not inactivated by 8, indicating that the inactivation of papain and cathepsin B by azapeptide esters is a specific process. Their ease of synthesis coupled with good solution stability suggests that azapeptide esters may be useful as active site titrants of cysteine proteinases and probes of their biological function in vivo.  相似文献   

16.
The aim of this study was to compare the effects of increased concentrations of MgADP, inorganic phosphate (Pi) and H+ ([MgADP], [Pi] and [H+], respectively) on the rate of relaxation in two different muscle types: skinned muscle fibres from the frog Rana temporaria and myofibrillar bundles from the giant Pacific acorn barnacle Balanus nubilus. Relaxation transients are produced by the photolysis of diazo-2 and are well fitted with a double exponential curve, giving two rate constants: k1 [5.6+/-0.1 s-1 for barnacle, n=30; 26.3+/-0.7 s-1 for frog, n=14 (mean+/-SEM)] and k2 [0.6+/-0.1 s-1 in barnacle, n=30; 10.4+/-1.0 s-1 in frog, n=14 (mean+/-SEM)], at 10 degrees C. Decreasing the pH by 0.5 pH units did not significantly affect k1 for barnacle relaxation [5.6+/-0.1 s-1 (mean+/-SEM), n=15] compared to the decrease in k1 of 40% seen in frog. Use of the Ca2+-sensitive fluorescent label acrylodan on barnacle wild-type troponin C demonstrated that decreasing the pH from 7.0 to 6.6 only alters the pCa50 value by 0.23 in the cuvette, while stopped-flow experiments with acrylodan revealed no significant change in koff from the labelled protein [322+/-32 s-1 at pH 7.0 and 381+/-24 s-1 (mean+/-SEM) at pH 6.6]. Increasing [MgADP] by 20 microM (50 microM added ADP) from control values of 50 microM in frog decreased k1 to 12.3+/-0.4 s-1 (mean+/-SEM, n=8), and at 400 microM MgADP, k1=9.6+/-0.1 s-1 (mean+/-SEM, n=12). In barnacle, 500 microM MgADP had a much smaller effect on k1 (4.0+/-0. 9 s-1, mean+/-SEM, n=8). Increasing the free [Pi] from the contaminant level of 0.36 mM to 1.9 mM slowed k1 by approximately 15% in barnacle [4.8+/-0.8 s-1, mean+/-SEM, n=7], compared to a approximately 30% reduction seen in frog. We conclude that the differences between barnacle and frog seen here are most probably due to different isomers of the contractile proteins, and that events underlying the crossbridge cycle are the same or similar. We interpret our results according to a model of crossbridge transitions during relaxation.  相似文献   

17.
The major source of amino acids for insect embryos are yolk proteins which accumulate in developing oocytes and are hydrolyzed during embryogenesis. Studies on Musca domestica embryogenesis indicated that a cathepsin B-like proteinase is responsible for yolk protein degradation (Ribolla et al., 1993). In this study, we report the purification of mature cathepsin and show that it is made up of a single 41 kDa polypeptide chain. The Musca domestica cathepsin NH2-terminal 11-residue sequence was determined (Ala-Pro-Lys-Tyr-Val-Asp-Tyr-Gly-Glu-Asn-Glu) and reveals homology with other cathepsins of the papain family. Experiments using serum anti-cathepsin show that the enzyme is stored in oocytes as a 55 kDa zymogen. The activation of the zymogen occurs in vitro only at low pH. In vitro activation in the presence of cysteine protease inhibitors is blocked at an intermediary polypeptide of 48 kDa. Kinetic studies of this activation process at pH 3.5 and 4.6 show that the zymogen is processed in a manner similar to that of pepsin (Foltmann, 1986) and papain (Vernet et al., 1991). We propose that Musca domestica cathepsin zymogen activation occurs in two steps. First, an intramolecular cleavage of the procathepsin polypeptide chain (55,000), induced by low pH gives rise to an intermediary polypeptide (48,000) which then undergoes autolysis to produce the mature enzyme (41,000).  相似文献   

18.
Methylglyoxal was demonstrated to be a substrate for the isozymes E1, E2 and E3 of human aldehyde dehydrogenase. Pyruvate was the product from the oxidation of methylglyoxal by the three isozymes. At pH 7.4 and 25 degrees C, the major and minor components of the E3 isozyme catalyzed the reaction with Vmax of 1.1 and 0.8 mumol NADH min-1 mg-1 protein, respectively, compared to 0.067 and 0.060 mumol NADH min-1 mg-1 protein for the E1 and E2 isozymes, respectively. The E2 isozyme had a K(m) for methylglyoxal of 8.6 microM, the lowest compared to 46 microM for E1 and 586 and 552 microM for the major and minor components of the E3 isozyme, respectively. Both components of the E3 isozyme showed substrate inhibition by methylglyoxal, with Ki values of 2.0 mM for the major component and 12 mM for the minor component at pH 9.0. Substrate inhibition by methylglyoxal was not observed with the E1 and E2 isozymes. Methylglyoxal strongly inhibited the glycolaldehyde activity of the E1 and E2 isozymes. Mixed-type models of inhibition were employed as an approach to calculate the inhibition constants, 44 and 10.6 microM for E1 and E2 isozymes, respectively.  相似文献   

19.
The effect of Ca2+ channel-acting drugs on bovine adrenal mitochondria Ca2+ movements was investigated. Mitochondrial Ca2+ uptake is performed by an energy-driven Ca2+ uniporter with a Km of 20.9 +/- 3.2 microM and Vmax of 148.1 +/- 7.2 nmol 45Ca2+ min-1 mg-1. Ca2+ release is performed through an Na+/Ca2+ antiporter with a Km for Na+ of 4.2 +/- 0.5 mM, a Vmax of 7.5 +/- 0.4 nmol 45Ca2+ min-1 mg-1, and a Hill coefficient of 1.4 +/- 0.2 Ca2+ efflux through the mitochondrial Na+/Ca2+ exchanger was inhibited by several dihydropyridines (nitrendipine, felodipine, nimodipine, (+)isradipine) and by the benzothiazepine diltiazem with similar potencies. In contrast, neither CGP 28392, Bay-K-8644, amlodipine, nor verapamil had any effect on Ca2+ efflux. Nitrendipine at 20 microM modified neither the Km nor the Hill coefficient for Na+, whereas the Vmax was reduced to 2.9 nmol 45Ca2+ min-1 mg-1, thus demonstrating noncompetitive modulation of the Na+/Ca2+ exchanger. None of the Ca2+ channel-acting drugs assayed at 100 microM affected Ca2+ influx through the uniporter. Ca2+ channel blockers inhibited the Na+/Ca2+ antiporter and displaced the specific binding of [3H]nitrendipine to intact mitochondria with Ki values similar to the IC50s obtained for the inhibition of the Ca2+ efflux. Ca2+ channel-acting drugs that did not inhibit the Na+/Ca2+ exchanger (amlodipine, CGP 28392, Bay-K-9644, and verapamil, at concentrations of 100 microM or higher) had no effect on [3H]nitrendipine binding. These results suggest that the adrenomedullary mitochondrial dihydropyridine receptor is associated with the Na+/Ca2+ exchanger.  相似文献   

20.
This study compared flow-sensitive magnetic resonance imaging with biplane transoesophageal echocardiography in combination with continuous wave Doppler from the suprasternal notch in patients with native coarctation or after surgical repair. Twenty patients (mean age 33 years, range 17-60) were investigated, of whom 15 had undergone surgery at mean age 13 years, range 5-43. Peak and mean flow in the ascending and descending aorta as well as coarctation peak velocity were determined with the magnetic resonance imaging phase contrast technique. Coarctation peak velocity was also measured by Doppler from the jugulum. Magnetic resonance imaging axial sections as well as biplane transoesophageal echocardiography were used to measure the smallest diameter of the constricted segment. Sixteen healthy volunteers, mean age 36 years, range 22-63, provided reference values for magnetic resonance imaging determined volume of flow in the aorta. Peak flow in the descending aorta was 9.2 +/- 3.7 l. min-1 (reference 13.0 +/- 2.5, P < 0.01) and mean flow 3.1 +/- 0.9 l. min-1 (reference 3.4 +/- 0.8, P > 0.05). The ratio of descending-to-ascending peak flow was 0.54 +/- 0.17 (reference 0.69 +/- 0.10, P < 0.01) and mean flow 0.68 +/- 0.15 (reference 0.69 +/- 0.08, P > 0.05). The coarctation velocity was slightly higher with Doppler than with magnetic resonance imaging (+0.24 +/- 0.44 m. s-1, 95% confidence interval +0.45 to +0.02 m. s-1, P = 0.05). The coarctation diameter was slightly larger with magnetic resonance imaging than with transoesophageal echocardiography (1.4 +/- 3.5 mm, 95% confidence interval +3.1 to -0.3 mm, P = 0.11). Both methods are suitable for the assessment and follow-up of coarctation of the aorta. Flow assessment with magnetic resonance imaging provides a hitherto unavailable measure with which to assess the severity of obstruction.  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