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 671 毫秒
1.
Vashishtha AK  West AH  Cook PF 《Biochemistry》2008,47(19):5417-5423
Kinetic studies were carried out for histidine-tagged saccharopine reductase from Saccharomyces cerevisiae at pH 7.0, suggesting a sequential mechanism with ordered addition of reduced nicotinamide adenine dinucleotide phosphate (NADPH) to the free enzyme followed by L-alpha-aminoadipate-delta-semialdehyde ( L-AASA) which adds in rapid equilibrium prior to l-glutamate in the forward reaction direction. In the reverse reaction direction, nicotinamide adenine dinucleotide phosphate (NADP) adds to the enzyme followed by addition of saccharopine. Product inhibition by NADP is competitive vs NADPH and noncompetitive vs alpha-AASA and L-glutamate, suggesting that the dinucleotide adds to the free enzyme prior to the aldehyde. Saccharopine is noncompetitive vs NADPH, alpha-AASA, and L-glutamate. In the direction of saccharopine oxidation, NADPH is competitive vs NADP and noncompetitive vs saccharopine, L-glutamate is noncompetitive vs both NADP and saccharopine, while L-AASA is noncompetitive vs saccharopine and uncompetitive vs NADP. The sequential mechanism is also corroborated by dead-end inhibition studies using analogues of AASA, L-glutamate, and saccharopine. 2-Amino-6-heptenoic acid was chosen as a dead-end analogue of L-AASA and is competitive vs AASA, uncompetitive vs NADPH, and noncompetitive vs L-glutamate. alpha-Ketoglutarate (alpha-Kg) serves as the dead-end analogue of L-glutamate and is competitive vs L-glutamate and uncompetitive vs L-AASA and NADPH. In the direction of saccharopine oxidation, N-oxalylglycine, L-pipecolic acid, L-leucine, alpha-ketoglutarate, glyoxylic acid, and L-ornithine were used as dead-end analogues of saccharopine and showed competitive inhibition vs saccharopine and uncompetitive inhibition vs NADP. The equilibrium constant for the reaction was measured at pH 7.0 by monitoring the change in absorbance of NADPH and is 200 M(-1). The value is in good agreement with the value determined using the Haldane relationship.  相似文献   

2.
The kinetic mechanism of the reaction catalyzed by glucose-6-phosphate dehydrogenase (EC 1.1.1.49) from Dicentrarchus labrax liver was examined using initial velocity studies, NADPH and glucosamine 6-phosphate inhibition and alternate coenzyme experiments. The results are consistent with a steady-state ordered sequential mechanism in which NADP+ binds first to the enzyme and NADPH is released last. Replots of NADPH inhibition show an uncommon parabolic pattern for this enzyme that has not been previously described. A kinetic model is proposed in agreement with our kinetic results and with previously published structural studies (Bautista et al. (1988) Biochem. Soc. Trans. 16, 903-904). The kinetic mechanism presented provides a possible explanation for the regulation of the enzyme by the [NADPH]/[NADP+] ratio.  相似文献   

3.
M F Carlier  D Pantaloni 《Biochemistry》1976,15(21):4703-4712
The binding of reduced nicotinamide adenine dinucleotide phosphate (NADPH) to nicotinamide adenine dinucleotide phosphate (NADP) dependent isocitrate dehydrogenase from beef liver cytoplasm was studied by several equilibrium techniques (ultracentrifugation, molecular sieving, ultrafiltration, fluorescence). Two binding sites (per dimeric enzyme molecule) were found with slightly different dissociation constants (0.5 and 0.12 muM) and fluorescence yields (7.7 and 6.3). A ternary complex was formed between enzyme, isocitrate, and NADPH, in which NADPH dissociation constant was 5 muM. On the contrary, no binding of NADPH to the enzyme took place in the presence of magnesium isocitrate. Dialysis experiments showed the existence of 1 NADP binding site/dimer, with a dissociation constant of 26 muM. When NADPH was present with the enzyme in the proportion of 1 molecule/dimer, the dissociation constant of NADP was decreased fourfold, reaching a value quantitatively comparable to the Michaelis constant. The kinetics of coenzyme binding was followed using the stopped-flow technique with fluorescence detection. NADPH binding to the enzyme occurred through one fast reaction (k1 = 20 muM-1 s-1). Dissociation of NADPH took place upon NADP binding; however, equilibrium as well as kinetic data were incompatible with a simple competition scheme. Dissociation of NADPH from the enzyme upon magnesium isocitrate binding was preceded by the formation of a transitory ternary complex in which the fluorescence of NADPH was only about 30% of that in the enzyme-NADPH complex. Then interaction between the conenzymes and the involvement of ternary complexes in the catalytic mechanism are discussed in relation with what is known about the regulatory role of the coenzyme (Carlier, M. F., and Pantaloni, D. (1976), Biochemistry, 15, 1761-1766).  相似文献   

4.
The chemical shifts of all the aromatic proton and anomeric proton resonances of NADP+, NADPH, and several structural analogues have been determined in their complexes with Lactobacillus casei dihydrofolate reductase by double-resonance (saturation transfer) experiments. The binding of NADP+ to the enzyme leads to large (0.9-1.6 ppm) downfield shifts of all the nicotinamide proton resonances and somewhat smaller upfield shifts of the adenine proton resonance. The latter signals show very similar chemical shifts in the binary and ternary complexes of NADP+ and the binary complexes of several other coenzymes, suggesting that the environment of the adenine ring is similar in all cases. In contrast, the nicotinamide proton resonances show much greater variability in position from one complex to another. The data show that the environments of the nicotinamide rings of NADP+, NADPH, and the thionicotinamide and acetylpyridine analogues of NADP+ in their binary complexes with the enzyme are quite markedly different from one another. Addition of folate or methotrexate to the binary complex has only modest effects on the nicotinamide ring of NADP+, but trimethoprim produces a substantial change in its environment. The dissociation rate constant of NADP+ from a number of complexes was also determined by saturation transfer.  相似文献   

5.
Steady state kinetic analysis at pH 7.0 of the reduction of DL-glyceraldehyde by pig muscle aldose reductase showed that the enzyme follows a sequential ordered mechanism with NADPH binding first. However, the "off constant" for NADP+ in the forward direction was 1 order of magnitude less than the kcat. Analysis of this anomaly by pre-steady state kinetics using stopped-flow fluorescence spectroscopy showed that this could be accounted for by isomerization of the enzyme-NADP+ complex and that the rate of isomerization is the rate-limiting step. The rate constant for this step was of the same order of magnitude as the kcat for the forward reaction. Fluorescence emission spectra of free and NADP(H)-bound enzyme suggested a conformational change upon binding of coenzyme. In the reverse direction (oxidation of glycerol) pre-steady state and steady state kinetic analyses were consistent with the rate-limiting step occurring before isomerization of the enzyme-NADPH complex. We conclude, therefore, that during the kinetic mechanism of the reduction of aldehydes by aldose reductase, a slow (kinetically detectable) conformational change in the enzyme occurs upon coenzyme binding. Since NADPH and NADP+ bind to the enzyme very tightly, this has implications for the targeting and binding of drugs that are aldose reductase inhibitors.  相似文献   

6.
A divalent cation electrode was used to measure the stability constants (association constants) for the magnesium and manganese complexes of the substrates for the NADP+-specific isocitrate dehydrogenase (EC 1.1.1.42) from pea stems. At an ionic strength of 26.5 mM and at pH 7.4 the stability constants for the Mg2+-isocitrate and Mg2+-NADP+ complexes were 0.85 +/- 0.2 and 0.43 +/- 0.04 mM-1 respectively and for the Mn2+-isocitrate and Mn2+-NADP+ complexes they were 1.25 +/- 0.07 and 0.75 +/- 0.09 mM-1 respectively. At the same ionic strength but at pH 6.0 the Mg2+-NADPH and Mn2+-NADPH complexes had stability constants of 0.95 +/- 0.23 and 1.79 +/- 0.34 mM-1 respectively. Oxalosuccinate and alpha-ketoglutarate do not form measureable complexes under these conditions. Saturation kinetics of the enzyme with respect to isocitrate and metal ions are consistent with the metal-isocitrate complex being the substrate for the enzyme. NADP+ binds to the enzyme in the free form. Saturation kinetics of NADPH and Mn2+ indicate that the metal-NADPH complex is the substrate in the reverse reaction. In contrast the pig heart enzyme appears to bind free NADPH and Mn2+. A scheme for the reaction mechanism is presented and the difference between the reversibility of the NAD+ and NADP+ enzyme is discussed in relation to the stability of the NADH and NADPH metal complexes.  相似文献   

7.
The binding of NADP+ to dihydrofolate reductase (EC 1.5.1.3) in the presence and absence of substrate analogs has been studied using 1H and 13C nuclear magnetic resonance (NMR). NADP+ binds strongly to the enzyme alone and in the presence of folate, aminopterin, and methotrexate with a stoichiometry of 1 mol of NADP+/mol of enzyme. In the 13C spectra of the binary and ternary complexes, separate signals were observed for the carboxamide carbon of free and bound [13CO]NADP+ (enriched 90% in 13C). The 13C signal of the NADP+-reductase complex is much broader than that in the ternary complex with methotrexate because of exchange line broadening on the binary complex signal. From the difference in line widths (17.5 +/- 3.0 Hz) an estimate of the dissociation rate constant of the binary complex has been obtained (55 +/- 10 sec-1). The dissociation rate of the NADP+-reductase complex is not the rate-limiting step in the overall reaction. In the various complexes studied large 13C chemical shifts were measured for bound [13CO]NADP+ relative to free NADP+ (upfield shifts of 1.6-4.3 ppm). The most likely origin of the bound shifts lies in the effects on the shieldings of electric fields from nearby charged groups. For the NADP+-reductase-folate system two 13C signals from bound NADP+ are observed indicating the presence of more than one form of the ternary complex. The IH spectra of the binary and ternary complexes confirm both the stoichiometry and the value of the dissociation rate constant obtained from the 13C experiments. Substantial changes in the IH spectrum of the protein were observed in the different complexes and these are distinct from those seen in the presence of NADPH.  相似文献   

8.
Kinetics of carbonyl reductase from human brain.   总被引:3,自引:1,他引:2       下载免费PDF全文
Initial-rate analysis of the carbonyl reductase-catalysed reduction of menadione by NADPH gave families of straight lines in double-reciprocal plots consistent with a sequential mechanism being obeyed. The fluorescence of NADPH was increased up to 7-fold with a concomitant shift of the emission maximum towards lower wavelength in the presence of carbonyl reductase, and both NADPH and NADP+ caused quenching of the enzyme fluorescence, indicating formation of a binary enzyme-coenzyme complex. Deuterium isotope effects on the apparent V/Km values decreased with increasing concentrations of menadione but were independent of the NADPH concentration. The results, together with data from product inhibition studies, are consistent with carbonyl reductase obeying a compulsory-order mechanism, NADPH binding first and NADP+ leaving last. No significant differences in the kinetic properties of three molecular forms of carbonyl reductase were detectable.  相似文献   

9.
6-Phosphogluconate dehydrogenase has been purified from human brain to a specific activity of 22.8 U/mg protein. The molecular weight was 90,000. At low ionic strengths enzyme activity increased, due to an increase in Vmax and a decrease in Km for 6-phosphogluconate, and activity subsequently decreased as the ionic strength was increased (above 0.12). Both 6-phosphogluconate and NADP+ provided good protection against thermal inactivation, with 6-phosphogluconate also providing considerable protection against loss of activity caused by p-chloromercuribenzoate and iodoacetamide. Initial velocity studies indicated the enzyme mechanism was sequential. NADPH was a competitive inhibitor with respect to NADP+, and the Ki values for this inhibition were dependent on the concentration of 6-phosphogluconate. Product inhibition by NADPH was noncompetitive when 6-phosphogluconate was the variable substrate, whereas inhibition by the products CO2 and ribulose 5-phosphogluconate and NADP+ were varied. In totality these data suggest that binding of substrates to the enzyme is random. CO2 and ribulose 5-phosphate are released from the enzyme in random order with NADPH as the last product released.  相似文献   

10.
The steady-state kinetics of the oxidative decarboxylation of 6-phosphogluconate catalysed by 6-phosphogluconate dehydrogenase from sheep liver in triethanolamine and phosphate buffers (pH 7.0) have been reinvestigated. In triethanolamine buffer the enzyme is inhibited by high NADP+ concentrations in the presence of low fixed concentrations of 6-phosphogluconate. Data are consistent with an asymmetric sequential mechanism in which NADP+ and 6-phosphogluconate bind randomly and product release is ordered. The pathway through the enzyme--6-phosphogluconate complex appears to be preferred in triethanolamine buffer. Pre-steady-state studies of the oxidative decarboxylation reaction at pH 6.0-8.0 show that hydride transfer is greater than 900 s-1. After the fast formation of NADPH in amounts equivalent to about half of the enzyme-active-centre concentration, the rate of NADPH formation is equal to the steady-state rate. Two possible interpretations are considered. Rapid fluorescence measurements of the displacement of NADPH from its complex with the enzyme at pH 6.0 and 7.0 indicate that the dissociation of NADPH is fast (greater than 800 s-1) and cannot be the rate-limiting step in oxidative decarboxylation. Coenzyme binding studies at equilibrium have been extended to include the determination of the dissociation constants for the binary complexes of enzyme with NADPH and NADP+ at pH 6.0-8.0 and the dissociation constant for NADPH in the ternary enzyme--6-phosphogluconate--NADPH complex in triethanolamine buffer, pH 7.0.  相似文献   

11.
Transhydrogenase couples the redox reaction between NADH and NADP+ to proton translocation across a membrane. The enzyme comprises three components; dI binds NAD(H), dIII binds NADP(H), and dII spans the membrane. The 1,4,5,6-tetrahydro analogue of NADH (designated H2NADH) bound to isolated dI from Rhodospirillum rubrum transhydrogenase with similar affinity to the physiological nucleotide. Binding of either NADH or H2NADH led to closure of the dI mobile loop. The 1,4,5,6-tetrahydro analogue of NADPH (H2NADPH) bound very tightly to isolated R. rubrum dIII, but the rate constant for dissociation was greater than that for NADPH. The replacement of NADP+ on dIII either with H2NADPH or with NADPH caused a similar set of chemical shift alterations, signifying an equivalent conformational change. Despite similar binding properties to the natural nucleotides, neither H2NADH nor H2NADPH could serve as a hydride donor in transhydrogenation reactions. Mixtures of dI and dIII form dI2dIII1 complexes. The nucleotide charge distribution of complexes loaded either with H2NADH and NADP+ or with NAD+ and H2NADPH should more closely mimic the ground states for forward and reverse hydride transfer, respectively, than previously studied dead-end species. Crystal structures of such complexes at 2.6 and 2.3 A resolution are described. A transition state for hydride transfer between dihydronicotinamide and nicotinamide derivatives determined in ab initio quantum mechanical calculations resembles the organization of nucleotides in the transhydrogenase active site in the crystal structure. Molecular dynamics simulations of the enzyme indicate that the (dihydro)nicotinamide rings remain close to a ground state for hydride transfer throughout a 1.4 ns trajectory.  相似文献   

12.
The steady-state kinetics of human erythrocyte glucose-6-phosphate dehydrogenase (D-glucose-6-phosphate: NADP+ 1-oxidoreductase, EC 1.1.1.49) dimers were studied by initial rate measurement. These experiments gave intersecting double-reciprocal plots suggesting a ternary complex mechanism with a Km for NADP and glucose 6-phosphate of 11 microM and 43 microM, respectively. These studies were combined with rate measurements in the presence of one product (NADPH), dead-end inhibitors, as well as alternative substrates. The inhibition by NADPH was found to be competitive with respect to both substrates. Alternate substrates experiments gave linear double-reciprocal plots over a wide range of substrate concentrations. The results suggest that the dimeric enzyme follows either a random or a Theorell-Chance mechanism.  相似文献   

13.
Ferredoxin:NADP+ oxidoreductase (ferredoxin: NADP+ reductase, EC 1.18.1.2) was shown to form a ternary complex with its substrates ferredoxin (Fd) and NADP(H), but the ternary complex was less stable than the separate binary complexes. Kd for oxidized binary Fd-ferredoxin NADP+ reductase complex was less than 50 nM; Kd(Fd) increased with NADP+ concentration, approaching 0.5-0.6 microM when the flavoprotein was saturated with NADP+ K(NADP+) also increased from about 14 microM to about 310 microM, on addition of excess Fd. The changes in Kd were consistent with negative cooperativity between the associations of Fd and NADP+ and with our unpublished observations which suggest that product dissociation is rate-limiting in the reaction mechanism. Similar interference in binding was observed in more reduced states; NADPH released much ferredoxin:NADP+ reductase from Fd-Sepharose whether the proteins were initially oxidized or reduced. Complexation between Fd and ferredoxin: NADP+ reductase was found to shield each center from paramagnetic probes; charge specificity suggested that the active sites of Fd and ferredoxin:NADP+ reductase were, respectively, negatively and positively charged.  相似文献   

14.
An effective means of relieving the toxicity of furan aldehydes, furfural (FFA) and 5-hydroxymethylfurfural (HMF), on fermenting organisms is essential for achieving efficient fermentation of lignocellulosic biomass to ethanol and other products. Ari1p, an aldehyde reductase from Saccharomyces cerevisiae, has been shown to mitigate the toxicity of FFA and HMF by catalyzing the NADPH-dependent conversion to corresponding alcohols, furfuryl alcohol (FFOH) and 5-hydroxymethylfurfuryl alcohol (HMFOH). At pH 7.0 and 25°C, purified Ari1p catalyzes the NADPH-dependent reduction of substrates with the following values (k(cat) (s(-1)), k(cat)/K(m) (s(-1)mM(-1)), K(m) (mM)): FFA (23.3, 1.82, 12.8), HMF (4.08, 0.173, 23.6), and dl-glyceraldehyde (2.40, 0.0650, 37.0). When acting on HMF and dl-glyceraldehyde, the enzyme operates through an equilibrium ordered kinetic mechanism. In the physiological direction of the reaction, NADPH binds first and NADP(+) dissociates from the enzyme last, demonstrated by k(cat) of HMF and dl-glyceraldehyde that are independent of [NADPH] and (K(ia)(NADPH)/k(cat)) that extrapolate to zero at saturating HMF or dl-glyceraldehyde concentration. Microscopic kinetic parameters were determined for the HMF reaction (HMF+NADPH?HMFOH+NADP(+)), by applying steady-state, presteady-state, kinetic isotope effects, and dynamic modeling methods. Release of products, HMFOH and NADP(+), is 84% rate limiting to k(cat) in the forward direction. Equilibrium constants, [NADP(+)][FFOH]/[NADPH][FFA][H(+)]=5600×10(7)M(-1) and [NADP(+)][HMFOH]/[NADPH][HMF][H(+)]=4200×10(7)M(-1), favor the physiological direction mirrored by the slowness of hydride transfer in the non-physiological direction, NADP(+)-dependent oxidation of alcohols (k(cat) (s(-1)), k(cat)/K(m) (s(-1)mM(-1)), K(m) (mM)): FFOH (0.221, 0.00158, 140) and HMFOH (0.0105, 0.000104, 101).  相似文献   

15.
p-Hydroxybenzoate hydroxylase (EC 1.14.13.2) from Pseudomonas fluorescens is a NADPH-dependent, FAD-containing monooxygenase catalyzing the hydroxylation of p-hydroxybenzoate to form 3,4-dihydroxybenzoate in the presence of NADPH and molecular oxygen. The mechanism of this three-substrate reaction was investigated in detail at pH 6.6, 4 degrees C, by steady state kinetics, stopped flow spectrophotometry, and equilibrium binding experiments. The initial velocity patterns are consistent with a ping-pong type mechanism which involves two ternary complexes between the enzyme and substrates. The first ternary complex is formed by random addition of p-hydroxybenzoate and NADPH to the enzyme, followed by the release of the first product (NADP+). The reduced enzyme . p-hydroxybenzoate complex now reacts with oxygen, the third substrate, to form the second ternary complex. The enzyme-bound p-hydroxybenzoate then reacts with the activated oxygen to give 3,4-dihydroxybenzoate which is released regenerating the oxidized enzyme for the next cycle. The binding of p-hydroxybenzoate to the oxidized enzyme to form a 1:1 complex causes large, characteristic spectral perturbations and fluorescence quenching. The dissociation constant for the enzyme . substrate complex was obtained by titrations in which absorbance and/or fluorescence quenching was measured. The binding constants of NADPH to the enzyme with and without p-hydroxybenzoate were determined kinetically by measuring the rate of reduction of the enzyme at different concentrations of NADPH. The reduction of the enzyme proceeds extremely slowly in the absence of p-hydroxybenzoate. The presence of the substrate causes a dramatic stimulation (140,000-fold) in the rate of enzyme reduction. The anaerobic reduction of the enzyme by NADPH in the presence of p-hydroxybenzoate produces a transient charge-transfer intermediate. On the basis of the proposed mechanism, the dissociation constants for p-hydroxybenzoate and NADPH as well as the Michaelis constants for all the three substrates were calculated from the initial velocity data. The agreement obtained between various kinetic parameters from the initial rate measurements and those calculated from the individual rate constants determined in rapid reactions, strongly supports the proposed mechanism for the p-hydroxybenzoate hydroxylase reaction.  相似文献   

16.
Interaction of ferredoxin-NADP+ reductase from Anabaena with its substrates   总被引:1,自引:0,他引:1  
The interaction of ferredoxin-NADP+ reductase from the cyanobacterium Anabaena variabilis with its substrates, NADP+ and ferredoxin, has been studied by difference absorption spectroscopy. Several structural analogs of NADP+ have been shown to form complexes the stabilities of which are strongly dependent on the ionic strength of the medium. In most cases the binding energy of these complexes and their difference absorption spectra are similar to those reported for the spinach enzyme. However, NADP+ perturbs the absorption spectra of the Anabaena and spinach enzymes in a different way. This difference has been shown to be related to the binding of the nicotinamide ring of NADP+ to the enzymes. These results are interpreted as being due to a different nicotinamide binding site in the two reductases. The enthalpic and entropic components of the Gibbs energy of formation of the NADP+ complex have been estimated. An increase in entropy on NADP+ binding seems to be the main source of stability for the complex. A shift of approximately 40 mV in the redox potential of the couple NADP+/NADPH has been observed to occur upon binding of NADP+ to the oxidized enzyme. This allows us to calculate the binding energy between the reductase and NADPH. The ability of the reductase, ferredoxin, and NADP+ to form a ternary complex indicates that the protein carrier binds to the reductase through a different site than that of the pyridine nucleotide.  相似文献   

17.
Kinetic and physicochemical properties of hamster liver diacetyl reductase have been examined. The results of kinetic studies on the reduction of diacetyl and NADPH to acetoin and NADP+ suggest that the reaction follows an Ordered Bi Bi mechanism in which NADPH binds first before diacetyl. The enzyme is a tetrameric glycoprotein of single subunits of a molecular weight of 23,500 with a sedimentation coefficient of 6.0S. The enzyme does not contain Zn, Cu, or Fe. The amino acid composition revealed an unusually low proportion of proline residues (0.9%). p-Chloromercuriphenylsulfonate and phenylglyoxal inactivated the enzyme, but the presence of NADPH prevented the loss of activity due to thiol and arginine modification. The enzyme transferred the pro 4S hydrogen atom of NADPH to the substrate and the binding of the enzyme to NADPH resulted in a red shift of the ultraviolet absorption spectrum of the cofactor.  相似文献   

18.
The kinetics of glucose-fructose oxidoreductase from Zymomonas mobilis   总被引:5,自引:0,他引:5  
Glucose-fructose oxidoreductase operates by a classic ping-pong mechanism with a single site for all substrates: glucose, fructose, gluconolactone and sorbitol. The Km values for these substrates were determined. The values of kcat are 200 s-1 and 0.8 s-1 for the forward and reverse directions respectively. The overall catalytic process consists of two half-reactions with alternate reduction of NADP+ and oxidation of NADPH tightly bound to the enzyme. Reduction of enzyme-NADP+ by glucose and oxidation of enzyme-NADPH by gluconolactone involve single first-order processes. The values of the rate constants at saturating substrate are 2100 s-1 and 8 s-1 respectively; deuterium isotope effects indicate that these are for the hydrogen transfer step. Oxidation of enzyme-NADPH by fructose is first order with a limiting rate constant of at least 430 s-1. The reaction of enzyme-NADP+ with sorbitol is biphasic, with rate constants for both phases less than 1 s-1. This behaviour is explained by a mechanism in which the slow cyclisation of the acyclic form of fructose follows its dissociation from the enzyme. The rate-determining steps for the overall reaction are probably dissociation of gluconolactone in the forward direction and hydrogen transfer from sorbitol to enzyme-bound NADP+ in the reverse direction.  相似文献   

19.
Initial rate, product inhibition, and alternate substrate studies of purified glucose 6-phosphate dehydrogenase of human blood platelets give results consistent with an Ordered BiBi reaction mechanism. NADP appears to be the first substrate to bind and NADPH the last product to be released. ADP and ATP inhibitions are both competitive with respect to glucose 6-phosphate. ADP inhibition is noncompetitive with respect to NADP. ATP inhibition with respect to NADP is complex and is interpreted to indicate that there are two ATP binding sites on the enzyme, one for which NADP can compete and one for which glucose 6-phosphate can compete.  相似文献   

20.
Succinic semialdehyde reductase, a NADP+-dependent enzyme, was purified from whole pig brain homogenates. The enzyme preparation migrates as a single protein and activity band on analytical gel electrophoresis. Succinic semialdehyde reductase (Mr 110,000) catalyzes the reduction of succinic semialdehyde to 4-hydroxybutyrate. The equilibrium constant of the reaction is Keq = 5.8 X 10(7) M-1 at pH 7 and 25 degrees C. The inhibition kinetic patterns obtained when 4-hydroxybutyrate or substrate analogs are used as inhibitors of the reaction catalyzed by the reductase are consistent with an ordered sequential mechanism, in which the coenzyme NADPH adds to the enzyme before the aldehyde substrate. A specific aldehyde reductase was also purified to homogeneity from brain mitochondria preparations. Its catalytic properties are identical to those of the enzyme isolated from whole brain homogenates. It is postulated that two enzymes, i.e. a NAD+-dependent dehydrogenase and a NADP+-dependent reductase, participate in the metabolism of succinic semialdehyde in the mitochondria matrix.  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