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.
NAD(+)-linked and NADP(+)-linked 3 alpha-hydroxysteroid dehydrogenases were purified to homogeneity from hamster liver cytosol. The two monomeric enzymes, although having similar molecular masses of 38,000, differed from each other in pI values, activation energy and heat stability. The two proteins also gave different fragmentation patterns by gel electrophoresis after digestion with protease. The NADP(+)-linked enzyme catalysed the oxidoreduction of various 3 alpha-hydroxysteroids, whereas the NAD(+)-linked enzyme oxidized the 3 alpha-hydroxy group of pregnanes and some bile acids, and the 17 beta-hydroxy group of testosterone and androstanes. The thermal stabilities of the 3 alpha- and 17 beta-hydroxysteroid dehydrogenase activities of the NAD(+)-linked enzyme were identical, and the two enzyme activities were inhibited by mixing 17 beta- and 3 alpha-hydroxysteroid substrates, respectively. Medroxyprogesterone acetate, hexoestrol and 3 beta-hydroxysteroids competitively inhibited 3 alpha- and 17 beta-hydroxysteroid dehydrogenase activities of the enzyme. These results show that hamster liver contains a 3 alpha(17 beta)-hydroxysteroid dehydrogenase structurally and functionally distinct from 3 alpha-hydroxysteroid dehydrogenase.  相似文献   

2.
Various kinetic approaches were carried out to investigate kinetic attributes for the dual coenzyme activities of mitochondrial aldehyde dehydrogenase from rat liver. The enzyme catalyses NAD(+)- and NADP(+)-dependent oxidations of ethanal by an ordered bi-bi mechanism with NAD(P)+ as the first reactant bound and NAD(P)H as the last product released. The two coenzymes presumably interact with the kinetically identical site. NAD+ forms the dynamic binary complex with the enzyme, while the enzyme-NAD(P)H complex formation is associated with conformation change(s). A stopped-flow burst of NAD(P)H formation, followed by a slower steady-state turnover, suggests that either the deacylation or the release of NAD(P)H is rate limiting. Although NADP+ is reduced by a faster burst rate, NAD+ is slightly favored as the coenzyme by virtue of its marginally faster turnover rate.  相似文献   

3.
The kinetics of 3alpha-hydroxysteroid : NAD oxidoreductase (EC 1.1.1.50) from Pseudomonas testosteroni (ATCC 11996) have been investigated. The kinetic analysis based on initial activity measurements and product inhibition studies, indicates that the addition of substrate to the enzyme and the release of products from it, follows an obligatory order (ordered bi bi mechanism). The ability of the enzyme to utilize the thionicotinamide analogue of NAD (sNAD) as cofactor has been investigated using various 3alpha-hydroxysteroids from both the C19, C21, and C24 series. The results show that the reaction velocity with sNAD as the cofactor is generally lower than with NAD. The decrease, however, varies considerably, being negligible with some steroids such as litocholic acid and deoxycholic acid and very pronounced with other such as tetrahydrocortisol and tetrahydrocortisone. The introduction of an 11beta-hydroxy or an 11-oxo group into the steroid molecule significantly reduces the ability of the enzyme to attack the 3alpha-hydroxy group. No such effect could be seen when the 11-hydroxy group was in the alpha-position. The results also indicate that, whereas NAD can serve as cofactor for both the monomeric and the dimeric forms of the enzyme, sNAD only acts as cofactor for the monomeric form. Thus sNAD is a valuable tool for the study of the reversible, concentration-depenedent monomeric-dimeric transition of the 3alpha-hydroxysteroid dehydrogenase.  相似文献   

4.
Rat ovarian 20α-hydroxysteroid dehydrogenase was purified 230-fold with a 48% recovery through a 3-step process involving hydrophobic, gel filtration and gree dye affinity chromatography. The purified enzyme was demonstrated to be a single polypeptide chain of Mr 36 000. Initial velocity studies of all four substrates in the forward and reverse reactions indicated a sequential mechanism for the enzyme. Product inhibition and dead-end inhibition studies with substrate analogs were consistent with an ordered bi-bi mechanism in which NADP is the first substrate bound to the enzyme and NADPH, the second product released, Several NADP analogs were demonstrated to function as coenzymes in the reaction catalyzed. The purified enzyme was denatured at moderate temperatures and the binding of NADP protected the enzyme against thermal denaturation.  相似文献   

5.
The dual nucleotide cofactor-specific enzyme, 3alpha-hydroxysteroid dehydrogenase (3alpha-HSD) from Pseudomonas sp. B-0831, is a member of the short-chain dehydrogenase/reductase (SDR) superfamily. Transient-phase kinetic studies using the fluorescence stopped-flow method were conducted with 3alpha-HSD to characterize the nucleotide binding mechanism. The binding of oxidized nucleotides, NAD(+), NADP(+) and nicotinic acid adenine dinucleotide (NAAD(+)), agreed well with a one-step mechanism, while that of reduced nucleotide, NADH, showed a two-step mechanism. This difference draws attention to previous characteristic findings on rat liver 3alpha-HSD, which is a member of the aldo-keto reductase (AKR) superfamily. Although functionally similar, AKRs are structurally different from SDRs. The dissociation rate constants associated with the enzyme-nucleotide complex formation were larger than the k(cat) values for either oxidation or reduction of substrates, indicating that the release of cofactors is not rate-limiting overall. It should also be noted that k(cat) for a substrate, cholic acid, with NADP(+) was only 6% of that with NAD(+), and no catalytic activity was detectable with NAAD(+), despite the similar binding affinities of nucleotides. These results suggest that a certain type of nucleotide can modulate nucleotide-binding mode and further the catalytic function of the enzyme.  相似文献   

6.
1. The binding of NAD(+) and NADP(+) to glutamate dehydrogenase has been studied in sodium phosphate buffer, pH7.0, by equilibrium dialysis. Approximate values for the dissociation constants are 0.47 and 2.5mm respectively. For NAD(+) the value agrees with that estimated from initial-rate results. 2. In the presence of the substrate analogue glutarate both coenzymes are bound more firmly, and there is one active centre per enzyme subunit. The binding results cannot be described in terms of independent and identical active centres, and binding is stronger at low coenzyme concentrations than at high concentrations. Either the six subunits of the oligomer are not identical or there are negative interactions between them in the binding of coenzymes in ternary complexes with glutarate. The latter explanation is favoured. 3. The binding studies support the conclusions drawn from earlier kinetic studies of the glutamate reaction. 4. ADP and GTP respectively decrease and increase the affinity of the enzyme for NAD(+) and NADP(+), in both the presence and absence of glutarate. The negative binding interactions in the presence of glutarate are abolished by ADP, which decreases the affinity for the coenzymes at low concentrations of the latter. 5. In the presence of glutarate, GTP and NAD(+) or NADP(+), the association of enzyme oligomers is prevented, and the solubility of the enzyme is decreased; the complex of enzyme and ligands readily crystallizes. 6. The results are discussed in relation to earlier kinetic studies.  相似文献   

7.
Initial velocity studies and product inhibition studies were conducted for the forward and reverse reactions of formaldehyde dehydrogenase (formaldehyde: NAD oxidoreductase, EC 1.2.1.1) isolated from a methanol-utilizing yeast Candida boidinii. The data were consistent with an ordered Bi-Bi mechanism for this reaction in which NAD+ is bound first to the enzyme and NADH released last. Kinetic studies indicated that the nucleoside phosphates ATP, ADP and AMP are competitive inhibitors with respect to NAD and noncompetitive inhibitors with respect to S-hydroxymethylglutathione. The inhibitions of the enzyme activity by ATP and ADP are greater at pH 6.0 and 6.5 than at neutral or alkaline pH values. The kinetic studies of formate dehydrogenase (formate:NAD oxidoreductase, EC 1.2.1.2) from the methanol grown C. boidinii suggested also an ordered Bi-Bi mechanism with NAD being the first substrate and NADH the last product. Formate dehydrogenase the last enzyme of the dissimilatory pathway of the methanol metabolism is also inhibited by adenosine phosphates. Since the intracellular concentrations of NADH and ATP are in the range of the Ki values for formaldehyde dehydrogenase and formate dehydrogenase the activities of these main enzymes of the dissimilatory pathway of methanol metabolism in this yeast may be regulated by these compounds.  相似文献   

8.
The NADP+-linked oestradiol-17 alpha dehydrogenase (EC 1.1.1.148) present in cell-free extracts of chicken liver was investigated with the aim of separating it from a closely related oestradiol-17 beta dehydrogenase (EC 1.1.1.62) found in the same subcellular fraction. However, its chromatographic behaviour on CM-cellulose and DEAE-cellulose was almost identical with that previously reported for the latter enzyme, including resolution into two peaks on the anion-exchanger. Both peaks contained oestradiol-17 alpha dehydrogenase and oestradiol-17 beta dehydrogenase activity. Further attempts to separate the putative enzymes by dye-ligand chromatography with the use of the dyes Procion Yellow, Reactive Red and Cibachron Blue linked to Sepharose were unsuccessful, and they behaved identically on affinity columns of adenosine 2',5'-bisphosphate-agarose and 17 beta-oestradiol 3-hemisuccinate bound to Sepharose. A previous report of partial separation on Sephadex G-200 was not confirmed. Slab gel electrophoresis of enzyme preparations after affinity chromatography on adenosine 2',5'-bisphosphate-agarose revealed multiple bands in systems containing sodium dodecyl sulphate, whereas analysis by rod gel electrophoresis gave two major and one minor bands that stained coincidently for oestradiol-17 alpha dehydrogenase, oestradiol-17 beta dehydrogenase, epitestosterone dehydrogenase and testosterone dehydrogenase activities. Isoelectric focusing gave four enzymically active peaks that each oxidized oestradiol-17 alpha and -17 beta. Apparent Km values for the two forms of oestradiol-17 alpha dehydrogenase obtained by DEAE-cellulose chromatography were 17 and 23 microM for oestradiol-17 alpha, and 8.7 and 11.0 microM for NADP+. Limited kinetic studies with oestradiol-17 alpha and -17 beta with the use of the mixed-substrate method showed that the total velocity was equal to the sum of the separate velocities. The active-site inhibitor-alkylating agent 17 beta-(1-oxoprop-2-ynyl)androst-4-en-3-one did not cause time- or temperature-dependent inhibition, in contrast with the reported case of the oestradiol-17 beta dehydrogenase and 20 alpha-hydroxysteroid dehydrogenase activities of the human placental oestradiol dehydrogenase. NADP+ appeared to afford some protection against inhibition. Investigation of substrate specificity with a limited range of steroids suggests that the enzyme(s) from chicken liver differs substantially from the oestradiol-17 beta dehydrogenase from human placenta, and although the evidence is not conclusive it suggests the existence of one enzyme.  相似文献   

9.
A monomeric 3 alpha-hydroxysteroid dehydrogenase with a molecular weight of 34,000 was purified to apparent homogeneity from mouse liver cytosol. The enzyme catalyzed the reversible oxidation of the 3 alpha-hydroxy group of C19-, C21-, and C24-steroids, reduced a variety of carbonyl compounds, and was inhibited by SH-reagents, synthetic estrogens, anti-inflammatory drugs, prostaglandins, and delta 4-3-ketosteroids. Although these properties are similar to those of the enzyme from rat liver cytosol, the mouse enzyme exhibited low dehydrogenase activity toward benzene dihydrodiol and some alicyclic alcohols, it showed a strict cofactor specificity for NADP(H), and high substrate inhibition was observed in the reverse reaction. In addition, dexamethasone, deoxycorticosterone, and medroxyprogesterone acetate inhibited the mouse enzyme competitively at low concentrations and noncompetitively at high concentrations, whereas hexestrol, indomethacin, and prostaglandin A1 were competitive inhibitors. Steady-state kinetic measurements in both directions indicated that the reaction proceeds through an ordered bi bi mechanism with the cofactors binding to the free enzyme. The 3-ketosteroid substrates inhibited the enzyme uncompetitively at elevated concentrations, suggesting that the substrates bind to the enzyme.NADPH complex and to the enzyme NADP+ complex.  相似文献   

10.
Ma H  Ratnam K  Penning TM 《Biochemistry》2000,39(1):102-109
Rat liver 3alpha-hydroxysteroid dehydrogenase (3alpha-HSD), an aldo-keto reductase, binds NADP(+) in an extended anti-conformation across an (alpha/beta)(8)-barrel. The orientation of the nicotinamide ring, which permits stereospecific transfer of the 4-pro-R hydride from NAD(P)H to substrate, is achieved by hydrogen bonds formed between the C3-carboxamide of the nicotinamide ring and Ser 166, Asn 167, and Gln 190 and by pi-stacking between this ring and Tyr 216. These residues were mutated to yield S166A, N167A, Q190A, and Y216S. In these mutants, K(d)(NADP(H)) increased by 2-11-fold but without a significant change in K(d)(NAD(H)). Steady-state kinetic parameters showed that K(m)(NADP)()+ increased 13-151-fold, and this was accompanied by comparable decreases in k(cat)/K(m)(NADP)()+. By contrast, K(m)(NAD)()+ increased 4-8-fold, but changes in k(cat)/K(m)(NAD)()+ were more dramatic and ranged from 23- to 930-fold. Corresponding changes in binding energies indicated that each residue contributed equally to the binding of NADP(H) in the ground and transition states. However, the same residues stabilized the binding of NAD(H) only in the transition state. These observations suggest that different modes of binding exist for NADP(H) and NAD(H). Importantly, these modes were revealed by mutating residues in the nicotinamide pocket indicating that direct interactions with the 2'-phosphate in the adenine mononucleotide is not the sole determinant of cofactor preference. The single mutations were unable to invert or racemize the stereochemistry of hydride transfer even though the nicotinamide pocket can accommodate both anti- and syn-conformers once the necessary hydrogen bonds are eliminated. When 4-pro-R-[(3)H]NADH was used to monitor incorporation into [(14)C]-5alpha-dihydrotestosterone, a decrease in the (3)H:(14)C ratio was observed in the mutants relative to wild-type enzyme reflecting a pronounced primary kinetic isotope effect. This observation coupled with the change in the binding energy for NAD(P)(H) in the transition state suggests that these mutants have altered the reaction trajectory for hydride transfer.  相似文献   

11.
Indanol dehydrogenase was purified to apparent homogeneity from monkey liver cytosol. The enzyme was a monomer with a molecular weight of 36,000 and pI of 8.7. The amino acid composition was determined. The enzyme oxidized alicyclic alcohols including transdihydrodiols of benzene and naphthalene in the presence of both NADP+ and NAD+, and reduced several xenobiotic carbonyl compounds in the presence of NADPH, the 4-pro-R hydrogen atom of which was transferred to the substrate. The results of fluorometric binding and kinetic studies are consistent with an ordered sequential mechanism with NADP+ binding first. The enzyme was inhibited competitively versus NADP+ and uncompetitively versus 1-indanol not only by chelating agents such as 1,10-phenanthroline and 2,2'-bipyridine but also by a nonchelating isomer, 4,4'-bipyridine, which suggests hydrophobic interaction of the aromatic compounds with the enzyme, which did not contain zinc. The enzyme was also inhibited by Cibacron blue dye, synthetic estrogens, and delta 4-3-ketosteroids. The inhibition by Cibacron blue was competitive versus NADP+ and noncompetitive versus 1-indanol, whereas those by hexestrol, medroxyprogesterone acetate, and progesterone were uncompetitive versus NADP+ and competitive versus 1-indanol, corraborating the ordered addition of the coenzyme prior to 1-indanol.  相似文献   

12.
A gene having high sequence homology (45-49%) with the glycerol-1-phosphate dehydrogenase gene from Methanobacterium thermoautotrophicum was cloned from the aerobic hyperthermophilic archaeon Aeropyrum pernix K1 (JCM 9820). This gene expressed in Escherichia coli with the pET vector system consists of 1113 nucleotides with an ATG initiation codon and a TAG termination codon. The molecular mass of the purified enzyme was estimated to be 38 kDa by SDS/PAGE and 72.4 kDa by gel column chromatography, indicating presence as a dimer. The optimum reaction temperature of this enzyme was observed to be 94-96 degrees C at near neutral pH. This enzyme was subjected to two-substrate kinetic analysis. The enzyme showed substrate specificity for NAD(P)H-dependent dihydroxyacetone phosphate reduction and NAD(+)-dependent glycerol-1-phosphate (Gro1P) oxidation. NADP(+)-dependent Gro1P oxidation was not observed with this enzyme. For the production of Gro1P in A. pernix cells, NADPH is the preferred coenzyme rather than NADH. Gro1P acted as a noncompetitive inhibitor against dihydroxyacetone phosphate and NAD(P)H. However, NAD(P)(+) acted as a competitive inhibitor against NAD(P)H and as a noncompetitive inhibitor against dihydroxyacetone phosphate. This kinetic data indicates that the catalytic reaction by glycerol- 1-phosphate dehydrogenase from A. pernix follows a ordered bi-bi mechanism.  相似文献   

13.
3-Hydroxyhexobarbital dehydrogenase, which catalyzes the reversible oxidation of 3-hydroxyhexobarbital to 3-oxohexobarbital, has been purified 470-fold from the soluble fraction of guinea pig liver with a yield of 47%. The specific activity of the purified enzyme is 9.4 units/mg of protein. Results of polyacrylamide gel disc electrophoresis and isoelectric focusing indicated that the purified enzyme preparation is a single and homogeneous protein. NADP+ served as preferred co-factor, but NAD+ is also utilized in the presence of phosphate ion. The guinea pig liver enzyme possessed a relatively narrow substrate specificity in comparison with the rabbit liver enzyme. It is very distinctive that guinea pig liver 3-hydroxyhexobarbital dehydrogenase catalyzes the dehydrogenation of 17beta-hydroxysteroids such as testosterone, 4-androstene-3beta,17beta-diol, 5alpha-androstane-3alpha,17beta-diol, 5alpha-androstane-3beta,17beta-diol, 5alpha-androstan-17beta-ol-3-one, and 5beta-androstane-3alpha,17beta-diol.  相似文献   

14.
The kinetic mechanism of two major monomeric 17 beta-hydroxysteroid dehydrogenases from mouse liver cytosol was studied at pH 7 in both directions with NADP(H) and three steroid substrates: testosterone, 5 beta-androstane-3 alpha, 17 17 beta-diol, and estradiol-17 beta. In each case the reaction mechanism of the two enzymes was sequential, and inhibition patterns by-products and dead-end inhibitors were consisted with an ordered bi bi mechanism with the coenzyme binding to the free enzyme, although there was difference in affinity and maximum velocity for the steroidal substrates between the two enzymes. Binding studies of the coenzyme and substrate indicate the binding of coenzyme to the free enzyme, in which 1 mol of NADPH binds to 1 mol of each monomeric enzyme. The 4-pro-R-hydrogen atom of NADPH was transferred to the alpha-face of the steroid molecule by the two enzymes.  相似文献   

15.
The locking-on strategy uses soluble analogues of the enzymes specific substrate to produce biospecific adsorption of individual NAD(P)(+)-dependent dehydrogenases on immobilized NAD(P)(+) derivatives, which is so selective that a single enzyme activity can be purified from crude cellular extracts in a single chromatographic step with yields approaching 100%. However, attempts to further develop and apply this strategy to the biospecific chromatographic purification of a range of NAD(P)(+)-dependent dehydrogenases revealed some anomalous chromatographic behavior and certain unexplained phenomenon. Much of this can be attributed to nonbiospecific interference effects. Identification and elimination of this interference is discussed in the present study focusing on bovine liver glutamate dehydrogenase (GDH; EC 1.4.1.3) as the "test" enzyme. Results further confirm the potential of the locking-on strategy for the rapid purification of NAD(P)(+)-dependent dehydrogenases and provide further insight into the parameters which should be considered during the development of a truly biospecific affinity chromatographic system based on the locking-on strategy. The kinetic mechanism of bovine liver GDH has been the topic of much controversy with some reports advocating a sequential ordered mechanism of substrate binding and others reporting a sequential random mechanism. Since the kinetic locking-on strategy is dependent on the target NAD(P)(+)-dependent dehydrogenase having an ordered sequential mechanism of substrate binding, the bioaffinity chromatographic behavior of bovine liver GDH using the locking-on tactic suggests that this enzyme has an ordered sequential mechanism of substrate binding under a variety of experimental conditions when NAD(+) is used as cofactor.  相似文献   

16.
The roles of particular amino acids in substrate and coenzyme binding and catalysis of glucose-6-phosphate dehydrogenase of Leuconostoc mesenteroides have been investigated by site-directed mutagenesis, kinetic analysis, and determination of binding constants. The enzyme from this species has functional dual NADP(+)/NAD(+) specificity. Previous investigations in our laboratories determined the three-dimensional structure. Kinetic studies showed an ordered mechanism for the NADP-linked reaction while the NAD-linked reaction is random. His-240 was identified as the catalytic base, and Arg-46 was identified as important for NADP(+) but not NAD(+) binding. Mutations have been selected on the basis of the three-dimensional structure. Kinetic studies of 14 mutant enzymes are reported and kinetic mechanisms are reported for 5 mutant enzymes. Fourteen substrate or coenzyme dissociation constants have been measured for 11 mutant enzymes. Roles of particular residues are inferred from k(cat), K(m), k(cat)/K(m), K(d), and changes in kinetic mechanism. Results for enzymes K182R, K182Q, K343R, and K343Q establish Lys-182 and Lys-343 as important in binding substrate both to free enzyme and during catalysis. Studies of mutant enzymes Y415F and Y179F showed no significant contribution for Tyr-415 to substrate binding and only a small contribution for Tyr-179. Changes in kinetics for T14A, Q47E, and R46A enzymes implicate these residues, to differing extents, in coenzyme binding and discrimination between NADP(+) and NAD(+). By the same measure, Lys-343 is also involved in defining coenzyme specificity. Decrease in k(cat) and k(cat)/K(m) for the D374Q mutant enzyme defines the way Asp-374, unique to L. mesenteroides G6PD, modulates stabilization of the enzyme during catalysis by its interaction with Lys-182. The greatly reduced k(cat) values of enzymes P149V and P149G indicate the importance of the cis conformation of Pro-149 in accessing the correct transition state.  相似文献   

17.
Three alcohol dehydrogenases have been identified in Acinetobacter calcoaceticus sp. strain HO1-N: an NAD(+)-dependent enzyme and two NADP(+)-dependent enzymes. One of the NADP(+)-dependent alcohol dehydrogenases was partially purified and was specific for long-chain substrates. With tetradecanol as substrate an apparent Km value of 5.2 microM was calculated. This enzyme has a pI of 4.5 and a molecular mass of 144 kDa. All three alcohol dehydrogenases were constitutively expressed. Three aldehyde dehydrogenases were also identified: an NAD(+)-dependent enzyme, an NADP(+)-dependent enzyme and one which was nucleotide independent. The NAD(+)-dependent enzyme represented only 2% of the total activity and was not studied further. The NADP(+)-dependent enzyme was strongly induced by growth of cells on alkanes and was associated with hydrocarbon vesicles. With tetradecanal as substrate an apparent Km value of 0.2 microM was calculated. The nucleotide-independent aldehyde dehydrogenase could use either Würster's Blue or phenazine methosulphate (PMS) as an artificial electron acceptor. This enzyme represents approximately 80% of the total long-chain aldehyde oxidizing activity within the cell when the enzymes were induced by growing the cells on hexadecane. It is particulate but can be solubilized using Triton X-100. The enzyme has an apparent Km of 0.36 mM for decanal.  相似文献   

18.
The kinetic mechanism of NADPH-dependent aldehyde reductase II and aldose reductase, purified from human placenta, has been studied using L-glucuronate and DL-glyceraldehyde as their respective substrates. For aldehyde reductase II, the initial velocity and product inhibition studies (using NADP and gulonate) indicate that the enzyme reaction sequence is ordered with NADPH binding to the free enzyme and NADP being the last product to be released. Inhibition patterns using menadione (an analog of the aldehydic substrate) and ATP-ribose (an analog of NADPH) are also consistent with a compulsory ordered reaction sequence. Isotope effects of deuterium-substituted NADPH (NADPD) also corroborate the above reaction scheme and indicate that hydride transfer is not the sole rate-limiting step in the reaction sequence. For aldose reductase, initial velocity patterns, product, and dead-end inhibition studies indicate a random binding pattern of the substrates and an ordered release of product; the coenzyme is released last. A steady-state random mechanism is also consistent with deuterium isotope effects of NADPD on the reaction sequence catalyzed by this enzyme. However, the hydride transfer step seems to be more rate determining for aldose reductase than for aldehyde reductase II.  相似文献   

19.
The enzyme 17β-hydroxysteroid dehydrogenase is required for the synthesis and 11β-hydroxysteroid dehydrogenase for the regulation of androgens in rat Leydig cells. This histochemical study describes ontogenetic changes in distribution and intensity of these enzymes in Leydig cells from postnatal day (pnd) 1–90. Using NAD or NADP as the cofactor, 17β-hydroxysteroid dehydrogenase (substrate: 5-androstene-3β, 17β-diol) peaks were observed on pnd 16 for fetal Leydig cells and on pnd 19 and 37 for adult Leydig cells. Between pnd 13 and 25 the fetal cells showed a higher intensity for the 17β-enzyme than the adult cells; more fetal Leydig cells were stained with NADP, whereas more adult cells were positive with NAD on pnd 13 and 16. A nearly identical distribution of 11β-hydroxysteroid dehydrogenase (substrate: corticosterone) was observed with NAD or NADP as the cofactor; the reaction was present from pnd 31 onwards, first in a few adult Leydig cells and later in almost all these cells homogeneously. The ontogenetic curves of the two enzymes show an inverse relationship. To conclude: (1) Generally, a stronger reaction for 17β-hydroxysteroid dehydrogenase is shown with NAD as cofactor than with NADP; using NADP, fetal Leydig cells show a stronger staining than adult Leydig cells. (2) The data possibly support the notion of a new isoform of 11β -hydroxysteroid dehydrogenase in addition to types 1 and 2.  相似文献   

20.
The kinetic mechanism of glucose dehydrogenase (EC 1.1.1.47) from Halobacterium salinarum was studied by initial velocity and product inhibition methods. The results suggest that both, in the forward and reverse direction, the reaction mechanism is of Bi Bi sequential ordered type involving formation of ternary complexes. NADP+ adds first and NADPH formed dissociates from the enzyme last. For the reverse direction, NADPH adds first and NADP+ leaves last. Product inhibition experiments indicate that (a), the coenzymes compete for the same site and form of the enzyme and (b), ternary abortive complexes of enzyme-NADP(+)-glucono-delta-lactone and enzyme-NADPH-glucose are formed. All the other inhibitions are noncompetitive.  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