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 171 毫秒
1.
Lactate dehydrogenase (EC 1.1.1.27) catalyzes the NAD-dependent oxidation to (oxalate) and reduction (to glycollate) of glyoxylate. The kinetics of this disproportionation are in accord with the usual reaction pathway of lactate dehydrogenase:substrate inhibition with appropriate pH dependence occurs; a steady state in the ratio of NADH to NAD+ is set up during the reaction, has the expected dependence on pH, and is independent of the initial glyoxylate, coenzyme, and enzyme concentration. At pH 7 the lactate dehydrogenase-NADH complex is about fivefold more likely to react with and reduce glyoxylate (at a concentration of 100 mm) than to dissociate to produce free NADH, and the ratio of the fraction of the enzyme-NADH complex which dissociates to the fraction which reacts with and reduces glyoxylate varies with glyoxylate concentration and with pH in a manner in agreement with the normal reaction pathway of the enzyme. With all concentrations of glyoxylate and over the pH range 7–9.6 both free (not enzyme bound) NAD+ and free NADH are formed in the steady state of the disproportionation. From these results it is apparent that lactate dehydrogenase, like alcohol dehydrogenase (EC 1.1.1.1), catalyzes a disproportionation within the bounds of its normal kinetic reaction pathway.  相似文献   

2.
The oxidation of indole-3-acetic acid (IAA) catalyzed by horseradish peroxidase (HRP) in the absence of added H2O2 was studied at pH 7.4 using spectral and kinetic approaches. Upon addition of a hundred-fold excess of IAA to HRP the native enzyme was rapidly transformed to compound II (HRP-II). HRP-II was the predominant catalytic enzyme species during the steady state. No compound III was observed. HRP-II was slowly transformed to the stable inactive verdohemo-protein, P-670. A precursor of P-670, so-called P-940 was not detected. After the cessation of IAA oxidation there was neither oxygen consumption nor P-670 formation; the remaining HRP-II was spontaneously reduced to native enzyme. Single exponential kinetics were observed in the steady state for IAA oxidation, oxygen consumption and P-670 formation yielding identical first order rate constants of about 6 . 10(4) s(-1). A comparison of the rate of IAA oxidation by HRP-II in the steady state and in the transient state indicated that more than 1 3 of the IAA was oxidized non-enzymatically during the steady state, confirming that a free radical chain reaction is involved in the peroxidase-catalyzed oxidation of IAA. IAA oxidation stopped before IAA was completely consumed, which cannot be ascribed to enzyme inactivation because 30-50% of the enzyme was still active after the end of the reaction. Instead, incomplete IAA oxidation is explained in terms of termination of the free radical chain reaction. Bimolecular rate constants of IAA oxidation by HRP-I and HRP-II determined under transient state conditions were (2.2 +/- 0.1) x 10(3) M(-1) s(-1) and (2.3 +/- 0.2) x 10(2) M(-1) s(-1).  相似文献   

3.
The steady state kinetic properties of a simple model for an enzyme catalyzed group transfer reaction between two substrates have been calculated. One substrate is assumed to bind slowly and the other rapidly to the enzyme. Apparent substrate inhibition or substrate activation by the rapidly binding substrate may result if the slowly binding substrate binds at unequal rates to the free enzyme and to the complex between the enzyme and the rapidly binding substrate. Competitive inhibition by each product with respect to its structurally analogous substrate is to be expected if both substrates are in rapid equilibrium with their enzyme-substrate complexes. This product inhibition pattern, however, may also be observed when one substrate binds slowly. Noncompetitive inhibition with respect to the rapidly binding substrate by its structurally analogous product may result if the slowly binding substrate binds more slowly to the enzyme-product complex than to the free enzyme. Inhibition by substrate analogs which are not products should follow the same rules as inhibition by products. Thus substrate analog inhibition experiments are not particularly informative. The form of inhibition by "transition state analog" inhibitors should reveal which substrate binds slowly. There is no sharp conceptual distinction between ordered and random "kinetic mechanisms". I therefore suggest that the use of these concepts should be abandoned.  相似文献   

4.
Light-mediated reduction of flavoproteins with flavins as catalysts   总被引:3,自引:0,他引:3  
It has been found that small amounts of free flavins greatly accelerate the photochemical reduction of flavoproteins both to the radical and fully reduced oxidation states. This catalytic effect has been shown to be due to the rapid photochemical reduction of the free flavin to its fully reduced state, followed by its reaction with the flavoprotein to yield flavoprotein radical and by its reaction with flavoprotein radical to yield fully reduced flavoprotein. Evidence is presented that the same route may occur with flavoproteins in the absence of added flavins. In this case the photoreduction is mediated by the small equilibrium concentration of free flavin coenzyme present in a flavorprotein solution. Hence, it is suggested that flavoprotein reduction with EDTA as photosubstrate does not involve an excited state of the holoprotein, nor contact of EDTA with the enzyme, but exchange of electrons between enzyme flavin and free reduced flavin.  相似文献   

5.
The kinetic behaviour of adsorptive enzyme systems with free and adsorbed enzyme forms in rapid equilibrium has been analysed. It has been shown that the dependences of enzymic reaction rate on substrate or “adsorptive effector” concentrations reveal the deviations from simple kinetic laws of Michaelis-Menten type (positive or negative kinetic co-operativity). Such kinetic anomalies should be observed when adsorption of the enzyme results in the changing catalytic properties and when the state of the equilibrium between free and bound enzyme forms depends on the presence of low molecular substances (substrates, coenzymes and various cellular metabolites). The physiological significance of adsorption-desorption processes for the enzyme activity regulation has been emphasized.  相似文献   

6.
The four half-transamination reactions [the pyridoxal form of Escherichia coli aspartate aminotransferase (AspAT) with aspartate or glutamate and the pyridoxamine form of the enzyme with oxalacetate or 2-oxoglutarate] were followed in a stopped-flow spectrometer by monitoring the absorbance change at either 333 or 358 nm. The reaction progress curves in all cases gave fits to a monophasic exponential process. Kinetic analyses of these reactions showed that each half-reaction is composed of the following three processes: (1) the rapid binding of an amino acid substrate to the pyridoxal form of the enzyme; (2) the rapid binding of the corresponding keto acid to the pyridoxamine form of the enzyme; (3) the rate-determining interconversion between the two complexes. This mechanism was supported by the findings that the equilibrium constants for half- and overall-transamination reactions and the steady-state kinetic constants (Km and kcat) agreed well with the predicted values on the basis of the above mechanism using pre-steady-state kinetic parameters. The significant primary kinetic isotope effect observed in the reaction with deuterated amino acid suggests that the withdrawal of the alpha-proton of the substrates is rate determining. The pyridoxal form of E. coli AspAT reacted with a variety of amino acids as substrates. The Gibbs free energy difference between the transition state and the unbound state (unbound enzyme plus free substrate), as calculated from the pre-steady-state kinetic parameters, showed a linear relationship with the accessible surface area of amino acid substrate bearing an uncharged side chain.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

7.
The excision of uracil bases from DNA is accomplished by the enzyme uracil DNA glycosylase (UNG). Recognition of uracil bases in free DNA is facilitated by uracil base pair dynamics, but it is not known whether this same mechanistic feature is relevant for detection and excision of uracil residues embedded in nucleosomes. Here we investigate this question using nucleosome core particles (NCPs) generated from Xenopus laevis histones and the high-affinity "Widom 601" positioning sequence. The reactivity of uracil residues in NCPs under steady-state multiple-turnover conditions was generally decreased compared to that of free 601 DNA, mostly because of anticipated steric effects of histones. However, some sites in NCPs had equal or even greater reactivity than free DNA, and the observed reactivities were not readily explained by simple steric considerations or by global DNA unwrapping models for nucleosome invasion. In particular, some reactive uracils were found in occluded positions, while some unreactive uracils were found in exposed positions. One feature of many exposed reactive sites is a wide DNA minor groove, which allows penetration of a key active site loop of the enzyme. In single-turnover kinetic measurements, multiphasic reaction kinetics were observed for several uracil sites, where each kinetic transient was independent of the UNG concentration. These kinetic measurements, and supporting structural analyses, support a mechanism in which some uracils are transiently exposed to UNG by local, rate-limiting nucleosome conformational dynamics, followed by rapid trapping of the exposed state by the enzyme. We present structural models and plausible reaction mechanisms for the reaction of UNG at three distinct uracil sites in the NCP.  相似文献   

8.
5-DeazaFAD bound to a hydrophobic site in apophotolyase and formed a stable reconstituted enzyme, similar to that observed with FAD. Although stoichiometric incorporation was observed, the flavin ring modification in 1-deazaFAD interfered with normal binding, decreased protein stability, and prevented formation of a stable flavin radical, unlike that observed with FAD. The results suggest that an important hydrogen bond is formed between the protein and N (1) in FAD, but not N (5), and that there is sufficient space at the normal flavin binding site near N (5) to accommodate an additional hydrogen but not near N (1). Catalytic activity was observed with enzyme containing 5-deazaFADH2 (42% of native enzyme) or 1-deazaFADH2 (11% of native enzyme) as its only chromophore, but no activity was observed with the corresponding oxidized flavins, similar to that observed with FAD and consistent with a mechanism where dimer cleavage is initiated by electron donation from excited reduced flavin to substrate. The protein environment in photolyase selectively enhanced photochemical reactivity in the fully reduced state, as evidenced by comparison with results obtained in model studies with the corresponding free flavins. Phosphorescence was observed with free or photolyase-bound 5-deazaFADH2, providing the first example of a flavin that exhibits phosphorescence in the fully reduced state. Formation of an enzyme-substrate complex resulted in a nearly identical extent of quenching of 5-deazaFADH2 phosphorescence (85.1%) and fluorescence (87.5%). The data are consistent with a mechanism involving exclusive reaction of substrate with the excited singlet state of 5-deazaFADH2, analogous to that proposed for FADH2 in native enzyme. Direct evidence for singlet-singlet energy transfer from enzyme-bound 5-deazaFADH2 to 5,10-CH(+)-H4folate was provided by the fact that pterin fluorescence was observed upon excitation of 5-deazaFADH2, accompanied by a decrease in 5-deazaFADH2 fluorescence. On the other hand, the fluorescence of enzyme-bound pterin was quenched by 5-deazaFADox, consistent with energy transfer from pterin to 5-deazaFADox. In each case, the spectral properties of the chromophores were consistent with the observed direction of energy transfer and indicated that transfer in the opposite direction was energetically unlikely. Unlike 5-deazaFAD, energy transfer from pterin to FAD is energetically feasible with FADH2 or FADox. The results indicate that the direction of flavin-pterin energy transfer at the active site of photolyase can be manipulated by changes in the flavin ring or redox state which alter the energy level of the flavin singlet.  相似文献   

9.
The wound-inducible lipoxygenase obtained from maize is one of the nontraditional lipoxygenases that possess dual positional specificity. In this paper, we provide our results on the determination and comparison of the kinetic constants of the maize lipoxygenase, with or without detergents in the steady state, and characterization of the dependence of the kinetic lag phase or initial burst, on pH, substrate, and detergent in the pre-steady state of the lipoxygenase reaction. The oxidation of linoleic acid showed a typical lag phase in the pre-steady state of the lipoxygenase reaction at pH 7.5 in the presence of 0.25% Tween-20 detergent. The reciprocal correlation between the induction period and the enzyme level indicated that this lag phenomenon was attributable to the slow oxidative activation of Fe (II) to Fe (III) at the active site of the enzyme as observed in other lipoxygenase reactions. Contrary to the lagging phenomenon observed at pH 7.5 in the presence of Tween-20, a unique initial burst was observed at pH 6.2 in the absence of detergents. To our knowledge, the initial burst in the oxidation of linoleic acid at pH 6.2 is the first observation in the lipoxygenase reaction. Kinetic constants (K(m) and k(cat) values) were largely dependent on the presence of detergent. An inverse correlation of the initial burst period with enzyme levels and interpretations on kinetic constants suggested that the observed initial burst in the oxidation of linoleic acid could be due to the availability of free fatty acids as substrates for binding with the lipoxygenase enzyme.  相似文献   

10.
Juers DH  Hakda S  Matthews BW  Huber RE 《Biochemistry》2003,42(46):13505-13511
The open-closed conformational switch in the active site of Escherichia coli beta-galactosidase was studied by X-ray crystallography and enzyme kinetics. Replacement of Gly794 by alanine causes the apoenzyme to adopt the closed rather than the open conformation. Binding of the competitive inhibitor isopropyl thio-beta-D-galactoside (IPTG) requires the mutant enzyme to adopt its less favored open conformation, weakening affinity relative to wild type. In contrast, transition-state inhibitors bind to the enzyme in the closed conformation, which is favored for the mutant, and display increased affinity relative to wild type. Changes in affinity suggest that the free energy difference between the closed and open forms is 1-2 kcal/mol. By favoring the closed conformation, the substitution moves the resting state of the enzyme along the reaction coordinate relative to the native enzyme and destabilizes the ground state relative to the first transition state. The result is that the rate constant for galactosylation is increased but degalactosylation is slower. The covalent intermediate may be better stabilized than the second transition state. The substitution also results in better binding of glucose to both the free and the galactosylated enzyme. However, transgalactosylation with glucose to produce allolactose (the inducer of the lac operon) is slower with the mutant than with the native enzyme. This suggests either that the glucose is misaligned for the reaction or that the galactosylated enzyme with glucose bound is stabilized relative to the transition state for transgalactosylation.  相似文献   

11.
The kinetic properties of highly purified preparations of sheep liver cytoplasmic aldehyde dehydrogenase (preparations that had been shown to be free from contamination with the corresponding mitochondrial enzyme) were investigated with both propionaldehyde and butyraldehyde as substrates. At low aldehyde concentrations, double-reciprocal plots with aldehyde as the variable substrate are linear, and the mechanism appears to be ordered, with NAD+ as the first substrate to bind. Stopped-flow experiments following absorbance and fluorescence changes show bursts of NADH production in the pre-steady state, but the observed course of reaction depends on the pre-mixing conditions. Pre-mixing enzyme with NAD+ activates the enzyme in the pre-steady state and we suggest that the reaction mechanism may involve isomeric enzyme--NAD+ complexes. High concentrations of aldehyde in steady-state experiments produce significant activation (about 3-fold) at high concentrations of NAD+, but inhibition at low concentrations of NAD+. Such behaviour may be explained by postulating the participation of an abortive complex in product release. Stopped-flow measurements at high aldehyde concentrations indicate that the mechanism of reaction under these conditions is complex.  相似文献   

12.
Mechanistic studies on the hydrolytic dehalogenation catalyzed by haloalkane dehalogenases are of importance for environmental and industrial applications. Here, Car-Parrinello (CP) and ONIOM hybrid quantum-mechanical/molecular mechanics (QM/MM) are used investigate the second reaction step of the catalytic cycle, which comprises a general base-catalyzed hydrolysis of an ester intermediate (EI) to alcohol and free enzyme. We focus on the enzyme LinB from Sphingomonas paucimobilis UT26, for which the X-ray structure at atomic resolution is available. In agreement with previous proposals, our calculations suggest that a histidine residue (His272), polarized by glutamate (Glu132), acts as a base, accepting a proton from the catalytic water molecule and transferring it to an alcoholate ion. The reaction proceeds through a metastable tetrahedral intermediate, which shows an easily reversed reaction to the EI. In the formation of the products, the protonated aspartic acid (Asp108) can easily adopt conformation of the relaxed state found in the free enzyme. The overall free energy barrier of the reaction calculated by potential of the mean force integration using CP-QM/MM calculations is equal to 19.5 +/- 2 kcal . mol(-1). The lowering of the energy barrier of catalyzed reaction with respect to the water reaction is caused by strong stabilization of the reaction intermediate and transition state and their preorganization by electrostatic field of the enzyme.  相似文献   

13.
Rat liver microsomal NADPH-cytochrome P-450 reductase was prepared free of detectable amounts of FMN by a new procedure based on the exchange of this flavin into apoflavodoxin. The resulting FMN-free reductase binds NADP in the oxidized state with the same affinity (Kd = 5 microM) and stoichiometry (1:1 molar ratio) as does the native enzyme. Both the native and FMN-free reductase catalyze rapid reduction of ferricyanide, but the ability to reduce th 5,6-benzoflavone-inducible form of the liver microsomal cytochrome P-450 (P-450LM4) is lost upon removal of FMN. The FMN-free enzyme was reconstituted with artificial flavins which, in the free state, have oxidation-reduction potentials ranging from -152 to -290 mV, including 5-carba-5-deaza-FMN and several FMN analogs with a halogen or sulfur substituent on the dimethylbenzene portion of the ring system. Enzyme reconstituted with 5-carba-5-deaza-FMN has catalytic properties which are not significantly different from those of the FMN-free reductase, and is unable to reduce P-450LM4. On the other hand, the ability to reduce P-450LM4 and the other FMN-dependent activities of the native reductase are restored by substitution of several other analogs for FMN, but the kinetics of P-450LM4 reduction, studied under anaerobic conditions by stopped flow spectrophotometry, are significantly altered. The oxidation-reduction behavior of enzyme reconstituted with 7-nor-7-Br-FMN is substantially different from that of the native enzyme, and less thermodynamic stabilization of the semiquinone is observed with this flavin analog. In contrast, the oxidation-reduction properties of enzyme containing 8-nor-8-mercapto-FMN are similar to those of the native enzyme, but the spectral properties are significantly different. As shown in a stopped flow experiment, reduction of this FMN analog precedes reduction of P-450LM4 when a complex of the flavoprotein and P-450LM4 is allowed to react with NADPH. Our experiments support a sequence of electron transfer in this enzyme system as follows: NADPH leads to FAD leads to FMN leads to P-450. We propose that the enzyme cycles between a le- and a 3e-reduced state during turnover and that electrons are donated to acceptors via the reaction, FMNH2 leads to FMNH ..  相似文献   

14.
J Aqvist  A Warshel 《Biochemistry》1989,28(11):4680-4689
Calculations of the free energy profile for the first two (rate-limiting) steps of the staphylococcal nuclease catalyzed reaction are reported. The calculations are based on the empirical valence bond method in combination with free energy perturbation molecular dynamics simulations. The calculated activation free energy is in good agreement with experimental kinetic data, and the catalytic effect of the enzyme is reproduced without any arbitrary adjustment of parameters. The enormous reduction of the activation barrier (relative to the reference reaction in water) appears to be largely associated with the strong electrostatic effect of the Ca2+ ion and the two arginine residues in the active site. This favorable electrostatic environment reduces the cost of the general-base catalysis step by almost 15 kcal/mol (by stabilizing the OH- nucleophile) and then stabilizes the developing negative charge on the 5'-phosphate group in the second step of the reaction by about 19 kcal/mol. The basic features of the originally postulated enzyme mechanism (Cotton et al., 1979) are found to be compatible with the observed activation free energy. However, the proposed modification of the mechanism (Sepersu et al., 1987), in which Arg 87 interacts only with the pentacoordinated transition state, is supported by the simulations. Further calculations on the D21E mutant also give results in good agreement with kinetic data.  相似文献   

15.
A general method is outlined that determines quantitatively the extent to which tight ligand binding to an enzyme active site is facilitated by the adoption of a stabler macromolecular conformation in the complex. The method therefore rejects the general assumption that competitive inhibitor binding to enzyme active sites involves only local (active site) interactions. The procedure involves comparing the unfolding transition state free energies of the free and complexed enzyme from physiological conditions. For the interaction of the transition state analog coformycin with bovine adenosine deaminase we observed that the binding free energy by the physiological enzyme was approximately 92% due to the assumption of a stabler enzyme conformation in the complex. The significance of these findings in terms of general enzyme catalysis is discussed.  相似文献   

16.
A general method is outlined that determines quantitatively the extent to which tight ligand binding to an enzyme active site is facilitated by the adoption of a stabler macromolecular conformation in the complex. The method therefore rejects the general assumption that competitive inhibitor binding to enzyme active sites involves only local (active site) interactions. The procedure involves comparing the unfolding transition state free energies of the free and complexed enzyme from physiological conditions. For the interaction of the transition state analog coformycin with bovine adenosine deaminase we observed that the binding free energy by the physiological enzyme was ~92% due to the assumption of a stabler enzyme conformation in the complex. The significance of these findings in terms of general enzyme catalysis is discussed.  相似文献   

17.
Proline racemase exists in two states, one of which binds and isomerizes L-proline and the other of which binds and isomerizes D-proline. In the enzyme-catalyzed racemization of proline at high substrate concentrations, the interconversion of the two forms of the free enzyme becomes rate limiting. The tracer perturbation method of Britton (1966, 1973) vividly demonstrates the kinetic importance of this enzyme interconversion under oversaturating conditions and allows an estimate of the rate constant for this reaction of 10(5) s-1. It is further shown that the enzyme is bound state saturated and the peak-switch concentration, Cp, is 125 mM. At substrate concentrations higher than 125 mM the enzyme becomes oversaturated, and the reaction rate is limited by the transition state for the interconversion of two forms of the free enzyme. It seems likely that the two free enzyme forms differ only in the protonation states of the acidic and basic groups at the active site.  相似文献   

18.
Lutropin-sensitive adenylate cyclase ((EC 4.6.1.1) ATP pyrophosphate-lyase (cyclizing)) in purified rat ovarian plasma membranes is stimulated by lutropin 2- to 3-fold in the absence, but 15- to 20-fold in the presence of GTP or p(NH)ppG. Following 10 to 15 min of incubation at 30 degrees C in the presence of lutropin, enzyme activity declined (50%) in the presence of GTP but not in the presence of p(NH)ppG. This desensitizing process induced by lutropin and GTP is not seen if NaF is also included in the incubation medium. The desensitized state of the enzyme persists at 4 degrees C in membranes washed free of the incubation medium. In this state the enzyme is characterized by: (i) a reduced response to lutropin even in the presence of p(NH)ppG; (ii) its response to NaF is not different from that of untreated enzyme; (iii) it reconverts to a fully responsive state following incubation (10 min, 30 degrees C) in GTP-free medium, a process accelerated by p(NH)ppG; (iv) the receptor content as well as the stability of the receptor.hormone complex does not differ from that of untreated fully responsive enzyme. It is proposed that desensitization results from a GTP-dependent, hormone-stimulated reaction that leads to impaired coupling of the enzyme system. The desensitized state induced is transient and may revert to a responsive one under specified conditions.  相似文献   

19.
The reaction mechanism of carboxypeptidase Y catalyzed reactions is investigated. Presteady state and steady state kinetic measurements are performed on the hydrolysis and aminolysis of an ester and an amide substrate. It is found that deacylation is the rate determining step in hydrolysis of the ester, pivalic acid 4-nitrophenol and acylation in that of the amide, succinyl-L-alanyl-L-alalyl-L-propyl-L-phenylalanine 4-nitroanilide.

The kinetic effects observed in the presence of a nucleophile, L-valine amide, where aminolysis occurs in parallel to the hydrolysis reaction are analysed in details. The results are described satisfactorily by a reaction scheme which involves the binding of the added nucleophile, (i) to the free enzyme, resulting in a simple competitive effect, and (ii) to the acyl-enzyme with the formation of a complex between the enzyme and the aminolysis product, the dissociation of which is rate determining. That scheme can account for both increases and decreases of kinetic parameter values as a function of the nucleophile concentration. There is no indication of binding of the nucleophile to the enzyme-substrate complex before acylation takes place.  相似文献   

20.
5-Ethylphenazine-lactate-dehydrogenase-NAD+ conjugate (EP(+)-LDH-NAD+) was prepared by linking poly(ethylene glycol)-bound 5-ethylphenazine and poly(ethylene glycol)-bound NAD+ to lactate dehydrogenase. The average number of the ethylphenazine moieties bound per molecule of enzyme subunit was 0.46, and that of the NAD+ moieties was 0.32. This conjugate is a semisynthetic enzyme having lactate oxidase activity using oxygen or 3-(4,5-dimethyl-2-thiazolyl)-2,5-diphenyl-2H-tetrazolium bromide (MTT) as an electron acceptor; to make such conjugates seems to be a general method for artificially converting a dehydrogenase into an oxidase. When the concentration of oxygen or MTT is varied, the oxidase activity fits the Michaelis-Menten equation with the following kinetic constants: for the reaction system with oxygen, the turnover number per subunit is 2.3 min-1 and Km for oxygen is 1.91 mM; and for the system with MTT, the turnover number is 0.25 min-1 and Km for MTT is 0.076 mM. At the initial steady state of the oxidase reaction, only 2.1% of the NAD+ moieties of the conjugate are in the free state (i.e. not bound in the coenzyme-binding site of the lactate dehydrogenase moiety) and the rest are hidden in the coenzyme site; almost all the NAD+ moieties are in the reduced state. The apparent intramolecular rate constant for the reaction between a free NADH moiety and an oxidized ethylphenazine moiety is 2.3 s-1 and 2.1 s-1 for the systems with oxygen and with MTT, respectively. The apparent effective concentration of the free NADH moiety for the ethylphenazine moiety is 5.5 microM and is much smaller than that (0.34 mM) of the ethylphenazine moiety for the free NADH moiety; this difference is due to the effect of hiding the NADH moiety in the binding site, as the hidden NADH moiety cannot react with the ethylphenazine moiety.  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