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 390 毫秒
1.
The kinetic properties of wheat germ sucrose phosphate synthetase, which catalyzes the reaction UDP-glucose + fructose 6-phosphate → UDP + sucrose 6-phosphate have been studied. A plot of the reciprocal initial velocity versus reciprocal substrate concentration gave a series of intersecting lines indicating a sequential mechanism. Product inhibition studies showed that UDP was competitive with UDP-glucose and noncompetitive with fructose 6-phosphate. A dead-end inhibitor, inorganic phosphate, was competitive with UDP-glucose and noncompetitive with fructose 6-phosphate. The results of initial velocity and product and dead-end inhibition studies suggested that the addition of substrates to the enzyme follows an ordered mechanism.  相似文献   

2.
The kinetic data on sugarcane (Saccharum spp. hybrids) sucrose synthase (SuSy, UDP-glucose: D-fructose 2-alpha-D-glucosyltransferase, EC 2.4.1.13) are limited. We characterized kinetically a SuSy activity partially purified from sugarcane variety N19 leaf roll tissue. Primary plot analysis and product inhibition studies showed that a compulsory order ternary complex mechanism is followed, with UDP binding first and UDP-glucose dissociating last from the enzyme. Product inhibition studies showed that UDP-glucose is a competitive inhibitor with respect to UDP and a mixed inhibitor with respect to sucrose. Fructose is a mixed inhibitor with regard to both sucrose and UDP. Kinetic constants are as follows: Km values (mm, +/- SE) were, for sucrose, 35.9 +/- 2.3; for UDP, 0.00191 +/- 0.00019; for UDP-glucose, 0.234 +/- 0.025 and for fructose, 6.49 +/- 0.61. values were, for sucrose, 227 mm; for UDP, 0.086 mm; for UDP-glucose, 0.104; and for fructose, 2.23 mm. Replacing estimated kinetic parameters of SuSy in a kinetic model of sucrose accumulation with experimentally determined parameters of the partially purified isoform had significant effects on model outputs, with a 41% increase in sucrose concentration and 7.5-fold reduction in fructose the most notable. Of the metabolites included in the model, fructose concentration was most affected by changes in SuSy activity: doubling and halving of SuSy activity reduced and increased the steady-state fructose concentration by about 42 and 140%, respectively. It is concluded that different isoforms of SuSy could have significant differential effects on metabolite concentrations in vivo, therefore impacting on metabolic regulation.  相似文献   

3.
Steady-state kinetic studies including initial velocity for mannitol oxidation and fructose reduction and product inhibition for mannitol oxidation using fructose and reduced nicotinamide adenine dinucleotide (NADH) are in accord with a reaction mechanism best described as ordered Bi-Bi with NAD+ and NADH designated as the first substrate, last product, respectively at pH 8.8. All replots of slopes and intercepts from product inhibition studies were linear. Dead-end inhibition studies using mannitol 1-phosphate gave slope-parabolic, intercept-linear noncompetitive inhibition for both NAD+ and mannitol as substrates. The dead-end inhibitor is capable of binding multiply to the E, EA, and EQ forms of the enzyme to an extent that is controlled by the concentration of substrates. The EQ complex is inferred to undergo a conformational change, E'Q equilibrium EQ, since (V1/E1) greater than (KiqV2)/(KqE1), and no evidence for dead-end complex formation with NADH can be adduced. This is interpreted to mean that the release of fructose from the central complex is faster than the isomerization of the E-NADH complex. When mannitol is saturating, the noncompetitive inhibition against NAD+, as the variable substrate, becomes parabolic uncompetitive. A replot of the slopes of the parabola against mannitol 1-phosphate remains concave upward. This situation could arise if the conformational change we infer in the EQ complex opens up additional sites on the protein which can interact with the dead-end inhibitor.  相似文献   

4.
A kinetic analysis of two homogeneous rat liver steroid (3 alpha-hydroxysteroid and 17 beta-hydroxysteroid) UDP-glucuronosyltransferases was conducted using bisubstrate kinetic analysis, product inhibition studies, and dead-end competitive inhibition studies. Double reciprocal plots of initial velocity versus substrate concentration, using bisubstrate kinetic analysis, gave a sequential mechanism. Product inhibition studies were compatible with either a rapid equilibrium, random-order kinetic mechanism or an ordered Theorell-Chance mechanism. Results of dead-end competitive inhibition studies excluded an ordered Theorell-Chance mechanism. The cumulative results are consistent with a rapid equilibrium random-order sequential kinetic mechanism for the glucuronidation of testosterone by purified 17 beta-hydroxysteroid UDP-glucuronosyltransferase and of androsterone by purified 3 alpha-hydroxysteroid UDP-glucuronosyltransferase.  相似文献   

5.
The kinetic mechanism of serine acetyltransferase from Haemophilus influenzae was studied in both reaction directions. The enzyme catalyzes the conversion of acetyl CoA and L-serine to O-acetyl-L-serine (OAS) and coenzyme A (CoASH). In the direction of L-serine acetylation, an equilibrium ordered mechanism is assigned at pH 6.5. The initial velocity pattern in the absence of added inhibitors is best described by a series of lines converging on the ordinate when L-serine is varied at different fixed levels of acetyl CoA. The initial velocity pattern at pH 7.5 is also intersecting, but the lines are nearly parallel. Product inhibition by OAS is noncompetitive against acetyl CoA, while it is uncompetitive against L-serine. Product inhibition by L-serine in the reverse reaction direction is noncompetitive with respect to both OAS and CoASH. Glycine and S-methyl-L-cysteine (SMC) were used as dead-end analogs of L-serine and OAS, respectively. Glycine is competitive versus L-serine and uncompetitive versus acetyl CoA, while SMC is competitive against OAS and uncompetitive against CoASH. Desulfo-CoA was used as a dead-end analog of both acetyl CoA and CoASH, and is competitive versus both substrates in the direction of L-serine acetylation; while it is competitive against CoASH and noncompetitive against OAS in the direction of CoASH acetylation. All of the above kinetic parameters are consistent with those predicted for an ordered mechanism at pH 6.5 with the exception of the uncompetitive inhibition by OAS vs. serine. The latter inhibition pattern suggests combination of OAS with the central E:acetyl CoA:serine complex. Cysteine is known to regulate its own biosynthesis at the level of SAT. As a dead-end inhibitor, L-cysteine is competitive against both substrates in both reaction directions. These results are discussed in terms of the mechanism of regulation.  相似文献   

6.
The effect of concentration of each substrate in the reaction catalyzed by sucrose synthetase isolated from sweet potato roots was determined. For the sucrose synthesizing reaction, UDP-glucose(ADP-glucose)+fructose→sucrose+UDP(ADP), the substrate saturation curves for UDP-glucose, ADP-glucose and fructose were hyperbolic in shape and the reaction was strongly inhibited by UDP competitively. On the other hand, the substrates for the reversal of sucrose synthetase reaction, sucrose+UDP(ADP)→UDP-glucose(ADP-glucose)+fructose, exhibited a sigmoidal shaped saturation curve which was deviated from the Michaelis-Menten equation. The plot of data according to the empirical Hill equation gives a values greater than 1.0 for every substrate examined in the latter case. In view of these experimental data, the major role of sucrose synthetase is postulated in that this enzyme is involved in the breakdown of sucrose in sweet potato root tissues instead of the sucrose synthesizing reaction. The molecular weight of the enzyme was determined to be about 540,000 by the Sephadex gel filtration chromatography.  相似文献   

7.
The kinetic mechanism of homoisocitrate dehydrogenase from Saccharomyces cerevisiae was determined using initial velocity studies in the absence and presence of product and dead end inhibitors in both reaction directions. Data suggest a steady state random kinetic mechanism. The dissociation constant of the Mg-homoisocitrate complex (MgHIc) was estimated to be 11 +/- 2 mM as measured using Mg2+ as a shift reagent. Initial velocity data indicate the MgHIc complex is the reactant in the direction of oxidative decarboxylation, while in the reverse reaction direction, the enzyme likely binds uncomplexed Mg2+ and alpha-ketoadipate. Curvature is observed in the double-reciprocal plots for product inhibition by NADH and the dead-end inhibition by 3-acetylpyridine adenine dinucleotide phosphate when MgHIc is the varied substrate. At low concentrations of MgHIc, the inhibition by both nucleotides is competitive, but as the MgHIc concentration increases, the inhibition changes to uncompetitive, consistent with a steady state random mechanism with preferred binding of MgHIc before NAD. Release of product is preferred and ordered with respect to CO2, alpha-ketoadipate, and NADH. Isocitrate is a slow substrate with a rate (V/E(t)) 216-fold slower than that measured with HIc. In contrast to HIc, the uncomplexed form of isocitrate and Mg2+ bind to the enzyme. The kinetic mechanism in the direction of oxidative decarboxylation of isocitrate, on the basis of initial velocity studies in the absence and presence of dead-end inhibitors, suggests random addition of NAD and isocitrate with Mg2+ binding before isocitrate in rapid equilibrium, and the mechanism approximates rapid equilibrium random. The Keq for the overall reaction measured directly using the change in NADH as a probe is 0.45 M.  相似文献   

8.
Data on initial velocity and isotope exchange at equilibrium suggest a nonclassical ping-pong mechanism for the dihydropyrimidine dehydrogenase from pig liver. Initial velocity patterns in the absence of inhibitors appeared parallel at low reactant concentration, with substrate inhibition by NADPH that is competitive with uracil and with substrate inhibition by uracil that is uncompetitive with NADPH. The Km values for both uracil (1 microM) and NADPH (7 microM) are low. As a result, it was difficult to determine whether the initial velocity pattern in the absence of added inhibitors was parallel. Thus, the pattern was redetermined in the presence of the dead-end inhibitor 2,6-dihydroxypyridine, which binds to both sites. This treatment effectively eliminates the inhibition by both substrates and increases their Km values, giving a strictly parallel pattern. Product and dead-end inhibition patterns are consistent with a mechanism in which NADPH reduces the enzyme at site 1 and electrons are transferred to site 2 to reduce uracil to dihydrouracil. The predicted mechanism is corroborated by exchange between [14C] NADP and NADPH as well as [14C]thymine and dihydrothymine in the absence of the other substrate-product pair.  相似文献   

9.
The kinetic mechanism of phosphofructokinase has been determined at pH 8 for native enzyme and pH 6.8 for an enzyme desensitized to allosteric modulation by diethylpyrocarbonate modification. In both cases, the mechanism is predominantly steady state ordered with MgATP binding first in the direction of fructose 6-phosphate (F6P) phosphorylation and rapid equilibrium random in the direction of MgADP phosphorylation. This is a unique kinetic mechanism for a phosphofructokinase. Product inhibition by MgADP is competitive versus MgATP and noncompetitive versus F6P while fructose 1,6-bisphosphate (FBP) is competitive versus fructose 6-phosphate and uncompetitive versus MgATP. The uncompetitive pattern obtained versus F6P is indicative of a dead-end E.MgATP.FBP complex. Fructose 6-phosphate is noncompetitive versus either FBP or MgADP. Dead-end inhibition by arabinose 5-phosphate or 2,5-anhydro-D-mannitol 6-phosphate is uncompetitive versus MgATP corroborating the ordered addition of MgATP prior to F6P. In the direction of MgADP phosphorylation, inhibition by anhydromannitol 1,6-bisphosphate is noncompetitive versus MgADP, while Mg-adenosine 5'(beta, gamma-methylene)triphosphate is noncompetitive versus FBP. Anhydromannitol 6-phosphate is a slow substrate, while anhydroglucitol 6-phosphate is not. This suggests that the enzyme exhibits beta-anomeric specificity.  相似文献   

10.
This work reports changes in sucrose synthase and invertase activities throughout endosperm development in wheat, together with the associated substrates and metabolites, sucrose, UDP, glucose, fructose and UDP-glucose. Throughout endosperm development, sucrose synthase had consistently higher activity than invertase and indeed invertase activity did not change appreciably. The observed variation in pattern and amounts of glucose and fructose present during the mid- and late stages of endosperm development confirmed the suggestion that invertase was not the preferred pathway of sucrose catabolism. Kinetic parameters for sucrose synthase were determined in crude extracts. Estimates of UDP and sucrose concentrations suggest that sucrose synthase is unlikely to achieve its potential maximum velocity. This limitation may however be overcome in part by the apparent excess catalytic activity measured during endosperm development.  相似文献   

11.
The properties of spinach leaf sucrose-phosphate synthetase (EC 2.4.1.14) and cytosolic fructose-1,6-bisphosphatase (EC 3.1.3.11) have been studied. These two enzymes have been considered to be important in the control of sucrose synthesis. Sucrose-phosphate synthetase from leaf tissue has not been studied in detail previously and we report a technique for purifying this enzyme 50-fold by chromatography on AH-Sepharose 4B. This method frees the enzyme from contaminants which interfere with assay procedures with little or no loss of activity. The partially purified enzyme has a Km for UDP-glucose of 7.1 mm and for fructose 6-phosphate of 0.8 mm. Fructose 1,6-bisphosphate, inorganic phosphate and UDP are strong inhibitors. The inhibition patterns of these suggest that the enzyme operates either by an ordered bi-bi or a Theorell-Chance mechanism. Partially purified cytosolic fructose-1,6-bisphosphatase is not only inhibited by AMP as previously reported, but is also inhibited by fructose 6-phosphate and UDP. From our observations, we conclude that sucrose biosynthesis is indeed controlled through these two enzymes and it appears that the rate of sucrose synthesis is largely dependent upon the supply of triose phosphate and ATP from the chloroplast.  相似文献   

12.
Porcine kidney betaine aldehyde dehydrogenase (EC 1.2.1.8) kinetic properties were determined at low substrate concentrations. The double-reciprocal plots of initial velocity versus substrate concentration are linear and intersect at the left of the 1/v axis and showed substrate inhibition with betaine aldehyde. Studies of inhibition by NADH and dead-end analogs showed that NADH is a mixed inhibitor against NAD(+) and betaine aldehyde. AMP is competitive with respect to NAD(+) and mixed with betaine aldehyde. Choline is competitive against betaine aldehyde and uncompetitive with respect to NAD(+). The kinetic behavior is consistent with an Iso-Ordered Bi-Bi Steady-State mechanism.  相似文献   

13.
Initial velocity, product inhibition, and substrate inhibition studies suggest that the endogenous lactate dehydrogenase activity of duck epsilon-crystallin follows an order Bi-Bi sequential mechanism. In the forward reaction (pyruvate reduction), substrate inhibition by pyruvate was uncompetitive with inhibition constant of 6.7 +/- 1.7 mM. In the reverse reaction (lactate oxidation), substrate inhibition by L-lactate was uncompetitive with inhibition constant of 158 +/- 25 mM. The cause of these inhibitions may be due to epsilon-crystallin-NAD(+)-pyruvate and epsilon-crystallin-NADH-L-lactate abortive ternary complex formation as suggested by the multiple inhibition studies. Pyruvate binds to free enzyme very poorly, with a very large dissociation constant. Bromopyruvate, fluoropyruvate, pyruvate methyl ester, and pyruvate ethyl ester are alternative substrates for pyruvate. 3-Acetylpyridine adenine dinucleotide, nicotinamide 1,N6-ethenoadenine dinucleotide, and nicotinamide hypoxanthine dinucleotide serve as alternative coenzymes for epsilon-crystallin. All the above alternative substrates or coenzymes showed an intersecting initial-velocity pattern conforming to the order Bi--Bi kinetic mechanism. Nicotinic acid adenine dinucleotide, thionicotinamide adenine dinucleotide, and 3-aminopyridine adenine dinucleotide acted as inhibitors for this enzymatic crystallin. The inhibitors were competitive versus NAD+ and noncompetitive versus L-lactate. alpha-NAD+ was a noncompetitive inhibitor with respect to the usual beta-NAD+. D-Lactate, tartronate, and oxamate were strong dead-end inhibitors for the lactate dehydrogenase activity of epsilon-crystallin. Both D-lactate and tartronate were competitive inhibitors versus L-lactate while oxamate was a competitive inhibitor versus pyruvate. We conclude that the structural requirements for the substrate and coenzyme of epsilon-crystallin are similar to those of other dehydrogenases and that the carboxamide carbonyl group of the nicotinamide moiety is important for the coenzyme activity.  相似文献   

14.
The kinetic mechanism of the hypothalamic NADPH-linked progesterone 5 alpha-reductase from female rats was determined to be equilibrium ordered sequential by initial velocity, product inhibition and dead-end inhibition studies. Analysis of the initial velocity data resulted in intersecting double reciprocal plots indicating a sequential mechanism (apparent Km (progesterone) = 95.4 +/- 4.5 nM; apparent Kia(NADPH) = 9.9 +/- 0.7 microM). The plot of 1/v vs 1/progesterone intersected on the ordinate which is consistent with an equilibrium ordered mechanism. Ordered addition of the substrates was also supported by product inhibition studies with NADP versus NADPH and NADP versus progesterone. NADP is a competitive inhibitor versus NADPH (apparent Kis = 4.3 +/- 1.3 microM) and a noncompetitive inhibitor versus progesterone (apparent Kis = 31.9 +/- 1.4 microM and apparent Kii = 145.4 +/- 15.5 microM). These inhibition patterns show that NADPH binds prior to progesterone. Taken together, these analyses indicate that the cofactor, NADPH, binds to the enzyme in rapid equilibrium and preferentially precedes the binding of progesterone.  相似文献   

15.
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   

16.
Initial velocity and product inhibition studies were carried out on UDP-glucose dehydrogenase (UDPglucose: NAD+ 6-oxidoreductase, EC 1.1.1.22) from beef liver to determine if the kinetics of the reaction are compatible with the established mechanism. An intersecting initial velocity pattern was observed with NAD+ as the variable substrate and UDPG as the changing fixed substrate. UDPglucuronic acid gave competitive inhibition of UDPG and non-competitive inhibition of NAD+. Inhibition by NADH gave complex patterns.Lineweaver-Burk plots of 1/upsilon versus 1/NAD+ at varied levels of NADH gave highly non-linear curves. At levels of NAD+ below 0.05 mM, non-competitive inhibition patterns were observed giving parabolic curves. Extrapolation to saturation with NAD+ showed NADH gave linear uncompetitive inhibition of UDPG if NAD+ was saturating. However, at levels of NAD+ above 0.10 mM, NADH became a competitive inhibitor of NAD+ (parabolic curves) and when NAD+ was saturating NADH gave no inhibition of UDPG. NADH was non-competitive versus UDPG when NAD+ was not saturating. These results are compatible with a mechanism in which UDPG binds first, followed by NAD+, which is reduced and released. A second mol of NAD+ is then bound, reduced, and released. The irreversible step in the reaction must occur after the release of the second mol of NADH but before the release of UDPglucuronic acid. This is apparently caused by the hydrolysis of a thiol ester between UDPglucoronic acid and the essential thiol group of the enzyme. Examination of rate equations indicated that this hydrolysis is the rate-limiting step in the overall reaction. The discontinuity in the velocities observed at high NAD+ concentrations is apparently caused by the binding of NAD+ in the active site after the release of the second mol of NADH, eliminating the NADH inhibition when NAD+ becomes saturating.  相似文献   

17.
Sucrose synthase catalyzes the reversible conversion of sucrose and UDP into fructose and UDP-glucose. In filamentous cyanobacteria, the sucrose cleavage direction plays a key physiological function in carbon metabolism, nitrogen fixation, and stress tolerance. In unicellular strains, the function of sucrose synthase has not been elucidated. We report a detailed biochemical characterization of sucrose synthase from Thermosynechococcus elongatus after the gene was artificially synthesized for optimal expression in Escherichia coli. The homogeneous recombinant sucrose synthase was highly specific for ADP as substrate, constituting the first one with this unique characteristic, and strongly suggesting an interaction between sucrose and glycogen metabolism.  相似文献   

18.
Kinetic studies of fructokinase I of pea seeds   总被引:3,自引:0,他引:3  
Fructokinase I of pea seeds has been purified to homogeneity and the enzyme shown to be monomeric, with a molecular weight of 72,000 +/- 4000. The reaction mechanism was investigated by means of initial velocity studies. Both substrates inhibited the enzyme; the inhibition caused by MgATP was linear-uncompetitive with respect to fructose whereas that caused by D-fructose was hyperbolic-noncompetitive against MgATP. The product D-fructose 6-phosphate caused hyperbolic-noncompetitive inhibition with respect to both substrates. MgADP caused noncompetitive inhibition, which gave intercept and slope replots that were linear with D-fructose but hyperbolic with MgATP. Free Mg2+ caused linear-uncompetitive inhibition when either substrate was varied. L-Sorbose and beta, gamma-methyleneadenosine 5'-triphosphate were used as analogs of D-fructose and MgATP, respectively. Inhibition experiments using these compounds indicated that substrate addition was steady-state ordered, with MgATP adding first. The product inhibition experiments were found to be consistent with a steady-state random release of products. The substrate inhibition caused by MgATP was most likely due to the formation of an enzyme-MgATP-product dead-end complex, whereas that caused by D-fructose was due to alternative pathways in the reaction mechanism. The inhibition caused by Mg2+ can be explained in terms of a dead-end complex with either a central complex or an enzyme-product complex.  相似文献   

19.
Radioactive sucrose, supplied through the cut base to Pisum sativum epicotyls, was transported to the growing apex (plumule and hook) and used there for the synthesis mainly of uridine diphosphoglucose (UDP- glucose), fructose and cell wall glucan. Enzyme extracts of the apical tissue contained sucrose synthetase activity which was freely reversible, i.e. formed UDP-glucose and fructose from sucrose (pH optimum = 6·6 for the cleavage reaction, Km for sucrose = 63 mM). Particulate fractions of the same tissue contained a β-glucan synthetase which utilized UDP-glucose for formation of alkali-soluble and -insoluble products (pH optimum = 8·4, Km for UDP-glucose = 1·9 mM). Values for Vmax and yields of these two synthetase activities were sufficient to account for observed rates of cellulose deposition during epicotyl growth (15–25 μg/hr/epicotyl). When soluble pea enzyme was supplied with sucrose and UDP at pH 6·6 and then the preparation was supplemented with particles bearing β-glucan synthetase at pH 8·4, the glucose moiety of sucrose was converted to glucan in vitro. The results indicate that it is feasible for these synthetases to co-operate in vivo to generate β-glucan for expanding cell walls.  相似文献   

20.
An analysis of the kinetic mechanism of the microsomal NADPH-linked progesterone 5 alpha-reductase obtained from female rat anterior pituitaries was performed. Initial velocity, product inhibition and dead-end inhibition studies indicate that the kinetic mechanism for the progesterone 5 alpha-reductase is equilibrium ordered sequential. Analysis of the initial velocity data resulted in intersecting double reciprocal plots suggesting a sequential mechanism [apparent Km(progesterone) = 88.2 +/- 8.2 nM; apparent Kia(NADPH) = 7.7 +/- 1.1 microM]. Furthermore, the plot of 1/v vs 1/progesterone intersected on the ordinate which is indicative of an equilibrium ordered mechanism. Additional support for ordered substrate binding was provided by the product inhibition studies with NADPH versus NADP and progesterone versus NADP. NADP is a competitive inhibitor versus NADPH (apparent Kis = 7.8 +/- 1.0 microM) and a noncompetitive inhibitor versus progesterone (apparent Kis = 9.85 +/- 2.1 microM and apparent Kii = 63.2 +/- 12.5 microM). These inhibition patterns suggest that NADPH binds prior to progesterone. In sum, these kinetic studies indicate that NADPH binds to the microsomal enzyme in rapid equilibrium and preferentially precedes the binding of progesterone.  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