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.
A method was developed to estimate the extended Michaelis constant and maximum velocity of a suicide substrate from the time-course of remaining enzyme activity with the use of simulation data calculated from the representative kinetic model for a suicide substrate proposed by Walsh et al. (Walsh, C., Cromartie, T., Marcotte, P. and Spencer, R. (1978) Methods Enzymol. 53, 437-448). For this purpose an analytical equation for the time-course of remaining enzyme activity, based on the suicide kinetic model, was derived by the steady-state method reported by Tatsunami et al. (Tatsunami, S., Yago, N. and Hosoe, M. (1981) Biochim. Biophys. Acta 662, 226-235). The accuracy of this analytical solution was proved by comparing the result with the exact solution obtained by numerical computation. A method was also developed to estimate the most important factor for a suicide substrate, the partition ratio, from the time-course of remaining enzyme activity.  相似文献   

2.
The reversibility of the unfolding-refolding transition of horse muscle phosphoglycerate kinase, induced by guanidine hydrochloride (Gdn X HCl), was studied using the regain of enzyme activity as a probe of the native structure. An irreversibility in the reactivation process was detected when the protein was incubated in a critical concentration of denaturant (0.7 +/- 0.1 M Gdn X HCl). This apparent irreversibility was observed for the unfolding process (N----D) as well as for the refolding process (D----N). The formation of the trough followed biphasic kinetics at 23 degrees C, the first phase obeying a first-order reaction corresponded to an isomerization of an intermediate; the second phase, protein-concentration-dependent, was suppressed by lowering the temperature to 4 degrees C. The structural properties of the inactive species were studied; all the beta structures were recovered, but about 29% of the helical structures remained unfolded, and two SH groups were buried. Simulated kinetics were compared with the experimental results and were used to extend the minimum folding scheme previously proposed from equilibrium and kinetic studies [Betton et al. (1984) Biochemistry 23, 6654-6661; Betton et al. (1985) Biochemistry 24, 4570-4577]. The intermediates trapped under these conditions were structured but devoid of catalytic activity. Taking into account the structural properties of these species, the nature of the interactions involved in their formation and stabilization is discussed.  相似文献   

3.
We have previously shown that superoxide radical anion (O2.-) reacts with hydroethidine (HE) to form a product that is distinctly different from ethidium (E+) (Zhao et al., Free Radic. Biol. Med. 34:1359; 2003). The structure of this product was recently determined as the 2-hydroxyethidium cation (2-OH-E+) (Zhao et al., Proc. Natl. Acad. Sci. USA 102:5727; 2005). In this study, using HPLC and mass spectrometry techniques, we show that 2-OH-E+ is formed from the reaction between HE and nitrosodisulfonate radical dianion (NDS) or Fremy's salt. The reaction kinetics and mechanism were determined using steady-state and time-resolved optical and EPR techniques. Within the first 50 ms, an intermediate was detected. Another intermediate absorbing strongly at 460 nm and weakly at 670 nm was detected within a second. The structure of this species was assigned to an imino quinone derivative of HE. The stoichiometry of the reaction indicates that two molecules of NDS were needed to oxidize a molecule of HE. We postulate that the first step of the reaction involves the hydrogen atom abstraction from HE to form an aminyl radical that reacts with another molecule of NDS to form an adduct that decomposes to an imino quinone derivative of HE. A similar mechanism has been proposed for the reaction between HE and O2.-. The reaction between HE and the Fremy's salt should provide a facile route for the synthesis of 2-OH-E+, a diagnostic marker product of the HE/O2.- reaction.  相似文献   

4.
Recent studies which estimated foot segment kinetic patterns were found to have inconclusive data on one hand, and did not dissociate the kinetics of the chopart and lisfranc joint. The current study aimed therefore at reproducing independent, recently published three-segment foot kinetic data (Study 1) and in a second stage expand the estimation towards a four-segment model (Study 2).Concerning the reproducibility study, two recently published three segment foot models (Bruening et al., 2014; Saraswat et al., 2014) were reproduced and kinetic parameters were incorporated in order to calculate joint moments and powers of paediatric cohorts during gait. Ground reaction forces were measured with an integrated force/pressure plate measurement set-up and a recently published proportionality scheme was applied to determine subarea total ground reaction forces. Regarding Study 2, moments and powers were estimated with respect to the Instituto Ortopedico Rizzoli four-segment model. The proportionality scheme was expanded in this study and the impact of joint centre location on kinetic data was evaluated.Findings related to Study 1 showed in general good agreement with the kinetic data published by Bruening et al. (2014). Contrarily, the peak ankle, midfoot and hallux powers published by Saraswat et al. (2014) are disputed. Findings of Study 2 revealed that the chopart joint encompasses both power absorption and generation, whereas the Lisfranc joint mainly contributes to power generation.The results highlights the necessity for further studies in the field of foot kinetic models and provides a first estimation of the kinetic behaviour of the Lisfranc joint.  相似文献   

5.
During the catalytic reaction of copper amine oxidase, one of the two prochiral hydrogen atoms at the C1 position of substrate amine is stereoselectively abstracted by a conserved Asp residue serving as a general base. Using stereospecifically deuterium-labeled enantiomers of 2-phenylethylamine, we previously showed that the pro-S alpha-proton is abstracted by the enzyme from Arthrobacter globiformis (AGAO) [Uchida, M., et al. (2003) Biosci. Biotechnol. Biochem. 67, 2664-2667]. More recently, we have also demonstrated that the pro-S selectivity of alpha-proton abstraction is fully retained even in the reaction of a mutant AGAO lacking the catalytic base [Chiu, Y.-C., et al. (2006) Biochemistry 45, 4105-4120]. On the basis of these findings, we have proposed that the stereoselectivity of alpha-proton abstraction is primarily determined by the conformation of the Schiff base intermediate formed between the substrate and the topa quinone cofactor (TPQ), stabilized by the binding of the distal part of the substrate to a hydrophobic pocket of the enzyme. In this conformation, the pro-S hydrogen atom to be abstracted is nearly perpendicular to the plane of the Schiff base-TPQ conjugate system, achieving the maximum overlap of sigma- and pi-orbitals. To further elucidate the stereochemical details, we have synthesized stereospecifically deuterium-labeled enantiomers of ethylamine, a very poor substrate for AGAO, in addition to those structurally related to the preferred substrate, 2-phenylethylamine. In marked contrast to the nearly complete pro-S selectivity of alpha-proton abstraction for most substrates that have been examined, the stereoselectivity for ethylamine decreased significantly to as little as 88%. The crystal structure of AGAO soaked with ethylamine showed very poor electron densities for the substrate Schiff base intermediate, showing that its conformation is not defined uniquely. Thus, the stereoselectivity of alpha-proton abstraction during the copper amine oxidase reaction is closely associated with the conformational flexibility of the substrate Schiff base intermediate.  相似文献   

6.
Acetohydroxyacid synthase (Ec 2.2.1.6) catalyses the thiamine diphosphate-dependent reaction between two molecules of pyruvate yielding 2-acetolactacte and CO2. The enzyme will also utilise hydroxypyruvate with a k(cat) value that is 12% of that observed with pyruvate. When hydroxypyruvate is the substrate, the enzyme undergoes progressive inactivation with kinetics that are characteristic of suicide inhibition. It is proposed that the dihydroxyethyl-thiamine diphosphate intermediate can expel a hydroxide ion forming an enol that rearranges to a bound acetyl group.  相似文献   

7.
The kinetics of microperoxidase-11 (MP-11) in the oxidation reaction of guaiacol (AH) by hydrogen peroxide was studied, taking into account the inactivation of enzyme during reaction by its suicide substrate, H2O2. Concentrations of substrates were so selected that: 1) the reaction was first-order in relation to benign substrate, AH and 2) high ratio of suicide substrate to the benign substrate, [H2O2] > [AH]. Validation and reliability of the obtained kinetic equations were evaluated in various nonlinear and linear forms. Fitting of experimental data into the obtained integrated equation showed a close match between the kinetic model and the experimental results. Indeed, a similar mechanism to horseradish peroxidase was found for the suicide-peroxide inactivation of MP-11. Kinetic parameters of inactivation including the intact activity of MP-11, alphai, and the apparent inactivation rate constant, ki, were obtained as 0.282 +/- 0.006 min(-1) and 0.497 +/- 0.013(-1) min at [H2O2] = 1.0 mM, 27 degrees C, phosphate buffer 5.0 mM, pH = 7.0. Results showed that inactivation of microperoxidase as a peroxidase model enzyme can occur even at low concentrations of hydrogen peroxide (0.4 mM).  相似文献   

8.
Studies [bachovchin, W. W., et al. (1978) Biochemistry 17, 2218] of the mechanism of inactivation of adenosylcobalamin-dependent diol dehydratase have led to the development of a general method to describe the kinetics of a reaction pathway containing a reservoir of mobile hydrogen. Analysis by this method of catalytic rate measurements for mixtures of 1,2-propanediol and 1,1-dideuterio-1,2-propanediol supports a mechanism involving an intermediate with three equivalent hydrogens, in which hydrogen transfer from this intermediate to product is the major rate-contributing step. Other results using tritium as a trace label [essenberg, M. K., et al. (1971) J. Am. Chem. Soc. 93, 1242] are considered in light of these deuterium isotope studies.  相似文献   

9.
The action of a potent tricyclic cholinesterase inhibitor ethopropazine on the hydrolysis of acetylthiocholine and butyrylthiocholine by purified horse serum butyrylcholinesterase (EC 3.1.1.8) was investigated at 25 and 37 degrees C. The enzyme activities were measured on a stopped-flow apparatus and the analysis of experimental data was done by applying a six-parameter model for substrate hydrolysis. The model, which was introduced to explain the kinetics of Drosophila melanogaster acetylcholinesterase [Stojan et al. (1998) FEBS Lett. 440, 85-88], is defined with two dissociation constants and four rate constants and can describe both cooperative phenomena, apparent activation at low substrate concentrations and substrate inhibition by excess of substrate. For the analysis of the data in the presence of ethopropazine at two temperatures, we have enlarged the reaction scheme to allow primarily its competition with the substrate at the peripheral site, but the competition at the acylation site was not excluded. The proposed reaction scheme revealed, upon analysis, competitive effects of ethopropazine at both sites; at 25 degrees C, three enzyme-inhibitor dissociation constants could be evaluated; at 37 degrees C, only two constants could be evaluated. Although the model considers both cooperative phenomena, it appears that decreased enzyme sensitivity at higher temperature, predominantly for the ligands at the peripheral binding site, makes the determination of some expected enzyme substrate and/or inhibitor complexes technically impossible. The same reason might also account for one of the paradoxes in cholinesterases: activities at 25 degrees C at low substrate concentrations are higher than at 37 degrees C. Positioning of ethopropazine in the active-site gorge by molecular dynamics simulations shows that A328, W82, D70, and Y332 amino acid residues stabilize binding of the inhibitor.  相似文献   

10.
The question of an interfacial versus a homogeneous reaction is carefully addressed for the enzymatic biphasic cleavage of mandelonitrile to benzaldehyde by Prunus amygdalus hydroxynitrile lyase (pa-Hnl) (Hickel et al. [1999] Biotechnol Bioeng 36:425-436). Experimental evidence, including 1) the reaction ceases when the interface is populated by previously adsorbed denatured pa-Hnl, 2) the reaction continues even after washout of the bulk enzyme from the aqueous phase, 3) highly nonpolar organic solvents initially promote fast reaction kinetics that relatively quickly decay to zero product production, and 4) the reaction rate is nonlinear in the bulk enzyme concentration, provide robust grounds for an interfacial reaction. We also model enzymatic mandelonitrile cleavage assuming a homogeneous aqueous-phase reaction. The homogeneous reaction scheme does not simultaneously account for the experimental observations of a linear dependence of the reaction rate on organic/water interfacial area, no dependence on the aqueous-phase volume, and a nonlinear dependence on pa-Hnl aqueous concentration. Further, simple calculations demonstrate that the homogeneous reaction rate is at least three orders of magnitude slower than those observed by Hickel et al. (1999). We again conclude that enzyme adsorbed at the organic solvent/water interface primarily catalyzes the biphasic mandelonitrile cleavage reaction.  相似文献   

11.
From calculations of a model reaction scheme for base-catalyzed RNA hydrolysis, a pentacoodinate dianionic intermediate 2a (Storer, et al., J. Am. Chem. Soc., 1991, 113, 5216-5219) as well as two transition states, TS1 and TS2, to the intermediate have been located by ab initio calculations at the 3-21G* level. Although the intermediate, which has the well depth on the order of kBT, is unlikely to be kinetically significant, the overall rate-limiting transition state structure TS2 obtained at 3-21G* level is very close to the corresponding structure at the STO-3G level; it has an extended P-O(5') bond breaking character. These gas-phase calculation results are used to qualitatively interpret mutagenesis results of Barnase and RNase T1 where water molecules are absent from the active site.  相似文献   

12.
Pseudomonas L-phenylalanine oxidase (deaminating and decarboxylating) contains two FAD molecules in one molecule of the enzyme (Koyama, H. (1983) J. Biochem. 93, 1313-1319). When the enzyme was mixed anaerobically with L-phenylalanine, beta-2-thienylalanine, L-tyrosine, or L-methionine, a spectral species (purple intermediate) with a broad absorption band around 540 nm was observed with each substrate, and decayed slowly. From the data on the overall reaction kinetics, the rate of the L-phenylalanine oxidase reaction was expressed as follows. e/v = e/Vm + A/[S] + B/[O2] where e represents the concentration of enzyme unit, v the rate of the overall reaction, Vm the maximum velocity, and A and B are constants. Furthermore, the reactions of the enzyme with beta-2-thienylalanine (mostly an oxygenase substrate) and L-methionine (an oxidase substrate) were analyzed by the "stopped flow" method. The following scheme for the mechanism of L-phenylalanine oxidase reaction with both substrates is proposed, based on the data obtained. (formula; see text) Where Eox represents the oxidized form of the enzyme unit, EoxS the enzyme unit (oxidized form)-substrate compound, X the purple intermediate with a characteristic broad absorption band around 540 nm, S the substrate and P the product.  相似文献   

13.
14.
In order to characterize the phosphoenzymes (EPs) formed from MgATP and CaATP as substrates, the effects of Mg2+ and Ca2+ outside SR vesicles on the hydrolysis rates of EPs were examined by using purified and unpurified Ca-ATPases of sarcoplasmic reticulum (SR) at low [gamma-32P]ATP (4-10 microM), 0.1 M KCl, pH 7.0, and 0 degrees C. When the phosphorylation reaction was stopped by adding an excess of EDTA over Ca and Mg, two components of EP, EPfast (rate constant, kfast = 15-20 min-1), and EPslow (kslow = 0.3-0.4 min-1), were recognized in the time course of EP decomposition. These two rates did not depend on the Ca2+ or Mg2+ concentration in the medium during the phosphorylation reaction, although the proportions of EPfast and EPslow essentially depended on the concentrations of MgATP and CaATP in the phosphorylation reaction medium. The proportion of EPfast increased with increasing [MgATP]/[CaATP] in the medium, whereas that of EPslow decreased. The rate of EPslow hydrolysis in the presence of excess EDTA was basically the same as that of EP formed from CaATP. These results suggest that EPfast and EPslow are derived from MgATP and CaATP, respectively, and EPfast is a reaction intermediate with Mg bound at the substrate site (MgEP), while the main EPslow is a reaction intermediate with Ca bound at the substrate site (CaEP) which is readily converted to metal-free EP by EDTA addition (Shigekawa et al., (1983) J. Biol. Chem. 258, 8698-8707). Mg2+ added outside SR vesicles stimulated the conversion of CaEP to MgEP and inhibited the hydrolysis of MgEP in the relatively high concentration range (K(Mg) = 7.9 mM). Ca2+ added outside SR vesicles stimulated the conversion of MgEP to CaEP and inhibited the conversion of CaEP to MgEP by Mg2+ addition. The Ca2+ outside SR vesicles did not essentially affect the hydrolysis of MgEP. These results suggest that the interconversion between MgEP and CaEP takes place during the reaction by exchange of the divalent cation on the substrate site. The following scheme is proposed. (formula: see text)  相似文献   

15.
The oxidative half-reaction of phenol hydroxylase has been studied by stopped-flow spectrophotometry. Three flavin-oxygen intermediates can be detected when the substrate is thiophenol, or m-NH2, m-OH, m-CH3, m-Cl, or p-OH phenol. Intermediate I, the flavin C(4a)-hydroperoxide, has an absorbance maximum at 380-390 nm and an extinction coefficient approximately 10,000 M-1 cm-1. Intermediate III, the flavin C(4a)-hydroxide, has an absorbance maximum at 365-375 nm and an extinction coefficient approximately 10,000 M-1 cm-1. Intermediate II has absorbance maxima of 350-390 nm and extinction coefficients of 10,000-16,000 M-1 cm-1 depending on the substrate. A Hammett plot of the logarithm of the rates of the oxygen transfer step, the conversion of intermediate I to intermediate II, gives a straight line with a slope -0.5. Fluoride ion is a product of the enzymatic reaction when 2,3,5,6-tetrafluorophenol is the substrate. These results are consistent with an electrophilic substitution mechanism for oxygen transfer. The conversions of I to II and II to III are acid-catalyzed. A kinetic isotope effect of 8 was measured for the conversion of II to III using deuterated resorcinol as substrate. The conversion of III to oxidized enzyme is base-catalyzed, suggesting that the reaction depends on the removal of the flavin N(5) proton. Product release occurs at the same time as the formation of intermediate III, or rapidly thereafter. The results are interpreted according to the ring-opened model of Entsch et al. (Entsch, B., Ballou, D. P., and Massey, V. (1976) J. Biol. Chem. 251, 2550-2563).  相似文献   

16.
17.
4-Nitrophenyl 2-(4-O-methyl-alpha-d-glucopyranuronosyl)-beta-d-xylopyranoside obtained on deesterification of 4-nitrophenyl 2-O-(methyl 4-O-methyl-alpha-d-glucopyranosyluronate)-beta-d-xylopyranoside (Hirsch et al., Carbohydr. Res. 310, 145-149, 1998) was found to be an excellent substrate for the measurement of hemicellulolytic alpha-glucuronidase activity. A new precise alpha-glucuronidase assay was developed by coupling the alpha-glucuronidase-catalyzed formation of 4-nitrophenyl beta-d-xylopyranoside with its efficient hydrolysis by beta-xylosidase. A recombinant strain of Saccharomyces cerevisiae, harboring and expressing the beta-xylosidase gene xlnD of Aspergillus niger under control of the alcohol dehydrogenase II promoter on a multicopy plasmid, was used as a source of beta-xylosidase. The activity values of beta-xylosidase in the assay required to achieve a steady-state rate of 4-nitrophenol formation shortly after starting the alpha-glucuronidase reaction were obtained both experimentally and by calculation using the kinetics of coupled enzyme reactions.  相似文献   

18.
1. The reactions of the adenosylcobalamin-dependent enzyme, ethanolamine ammonia-lyase, with the 'good' and 'relatively poor' substrates 2-aminoethanol and (S)-2-aminopropanol respectively, under conditions of saturation with substrate were investigated by rapid freezing in conjunction with electron paramagnetic resonance (e.p.r.) spectroscopy and by stopped-flow spectrophotometry. 2. In disagreement with earlier reports [Babior et al. (1972) J. Biol. Chem. 247, 4389-4392], it was found that the reaction of 2-aminoethanol gave an e.p.r. signal observed in rapid freezing experiments characteristic of a coupled Co(II)-free radical system. This signal was similar to, though not identical with, that obtained with (S)-2-aminopropanol. The steady-state level of the signal with 2-aminoethanol as substrate was 0.56 of that attained with (S)-2-aminopropanol. 3. The results of these e.p.r. experiments were shown to be consistent with stopped-flow data obtained under closely similar reaction conditions, the latter indicating a corresponding ratio of 0.64. The results also are consistent with those of a rapid wavelength scanning, stopped-flow spectrophotometric study [Hollaway et al. (1978) Eur. J. Biochem. 82, 143-154].  相似文献   

19.
Whereas sitosterol and 24(28)-methylene cycloartanol were competitive inhibitors (with Ki = 26 microM and 14 microM, respectively), 24(R,S)-25-epiminolanosterol was found to be a potent non-competitive inhibitor (Ki = 3.0 nM) of the S-adenosyl-L-methionine-C-24 methyl transferase from sunflower embryos. Because the ground state analog, 24(R,S)-oxidolanosterol, failed to inhibit the catalysis and 25-azalanosterol inhibited the catalysis with a Ki of 30 nM we conclude that the aziridine functions in a manner similar to the azasteriod (Rahier, A., et al., J. Biol. Chem. (1984) 259, 15215) as a transition state analog mimicking the carbonium intermediate found in the normal transmethylation reaction. Additionally, we observed that the aziridine inhibited cycloartenol metabolism (the preferred substrate for transmethylation) in cultured sunflower cells and cell growth.  相似文献   

20.
Wu G  Kulmacz RJ  Tsai AL 《Biochemistry》2003,42(46):13772-13777
The peroxidase and cyclooxygenase activities of prostaglandin H synthase-1 (PGHS-1) both become irreversibly inactivated during reaction with peroxide. Sequential stopped-flow absorbance measurements with a chromogenic peroxidase cosubstrate previously were used to evaluate the kinetics of peroxidase inactivation during reaction of PGHS-1 with peroxide [Wu, G., et al. (1999) J. Biol. Chem. 274, 9231-7]. This approach has now been adapted to use a chromogenic cyclooxygenase substrate to analyze the detailed kinetics of cyclooxygenase inactivation during reaction of PGHS-1 with several hydroperoxides. In the absence of added reducing cosubstrates, which maximizes the levels of oxidized enzyme intermediates expected to lead to inactivation, cyclooxygenase activity was lost as fast as, or somewhat faster than, peroxidase activity. Cyclooxygenase inactivation kinetics appeared to be sensitive to the structure of the peroxide used. The addition of reducing cosubstrate during reaction of PGHS-1 with peroxide protected the peroxidase activity to a much greater degree than the cyclooxygenase activity. The results suggest a new concept of PGHS inactivation: that distinct damage can occur at the two active sites during side reactions of Intermediate II, which forms during reaction of PGHS with peroxide and which contains two oxidants, a ferryl heme in the peroxidase site, and a tyrosyl free radical in the cyclooxygenase site.  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