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.
We have applied the soluble pyridine nucleotide transhydrogenase of Pseudomonas fluorescens to a cell-free system for the regeneration of the nicotinamide cofactors NAD and NADP in the biological production of the important semisynthetic opiate drug hydromorphone. The original recombinant whole-cell system suffered from cofactor depletion resulting from the action of an NADP(+)-dependent morphine dehydrogenase and an NADH-dependent morphinone reductase. By applying a soluble pyridine nucleotide transhydrogenase, which can transfer reducing equivalents between NAD and NADP, we demonstrate with a cell-free system that efficient cofactor cycling in the presence of catalytic amounts of cofactors occurs, resulting in high yields of hydromorphone. The ratio of morphine dehydrogenase, morphinone reductase, and soluble pyridine nucleotide transhydrogenase is critical for diminishing the production of the unwanted by-product dihydromorphine and for optimum hydromorphone yields. Application of the soluble pyridine nucleotide transhydrogenase to the whole-cell system resulted in an improved biocatalyst with an extended lifetime. These results demonstrate the usefulness of the soluble pyridine nucleotide transhydrogenase and its wider application as a tool in metabolic engineering and biocatalysis.  相似文献   

2.
吡啶核苷酸转氢酶(pyridine nucleotide transhydrogenases)是能直接催化NADP(H)与NAD(H)之间氢负离子可逆转移的氧化还原酶, 主要调控分解代谢和合成代谢过程中NADH与NADPH之间的动态平衡. 膜结合吡啶核苷酸转氢酶(TH)是ATP依赖性跨膜蛋白, 由2个亚基构成, 每个亚基包含dⅠ、dⅡ和dⅢ三个结构域. 在TH结合可变催化机制中, 氢负离子的转移总是与质子转移相偶联. 可溶性吡啶核苷酸转氢酶(STH)是非能量依赖性的黄素蛋白, 以可溶性多聚体形式存在. 目前认为,很多因活性氧自由基异常增多而引起的线粒体疾病都与TH的活性有关, 包括糖尿病、癌症、神经退行性疾病及心血管疾病等. TH分子机制的研究将有助于揭示这些线粒体疾病的致病机理以及为其诊断和基因治疗提供分子依据. STH作用机理的研究及其在辅酶再生系统中的应用, 将会推动代谢工程和工业生物催化过程的进一步发展.  相似文献   

3.
Morphine 6-dehydrogenase, which catalyzes the dehydrogenation of morphine to morphinone, has been purified about 440-fold from the soluble fraction of guinea pig liver with a yield of 38%. The purified enzyme was a homogeneous protein on polyacrylamide gel disc electrophoresis and isoelectric focusing. The molecular weight and isoelectric point of the enzyme were 29,000 and 7.6, respectively. The enzyme utilizes both NAD and NADP as a cofactor, and the Km values were 0.12 mM for NAD and 0.42 mM for NADP. The Vmax values for morphine were 588 milliunits/mg of protein (with NAD) and 1600 milliunits/mg of protein (with NADP). The Km values for morphine were 0.12 mM (with NAD) and 0.49 mM (with NADP). The enzyme also exhibited activity for morphine-related compounds: nalorphine, normorphine, codeine, and ethylmorphine; however, 7,8-saturated congeners such as dihydromorphine and dihydrocodeine were poor substrates. The enzyme was inactivated by removal of 2-mercaptoethanol from the enzyme solution. The inactivated enzyme was rapidly recovered by the addition of 2-mercaptoethanol. Phenylarsine oxide and CdCl2 (dithiol modifiers) inhibited competitively toward cofactor binding and noncompetitively toward morphine binding. These results suggest that the enzyme possesses the essential thiol groups, probably vicinal dithiol, at or near the cofactor-binding site. Using the partially purified enzyme, 8-(2-hydroxyethylthio)dihydromorphinone was isolated as the product and identified by UV, mass, and NMR spectra. It was confirmed that morphinone proposed as the dehydrogenation product was nonenzymatically and covalently bound to 2-mercaptoethanol. Accordingly, the isolated morphinone-2-mercaptoethanol conjugate must be formed by two steps: enzymatic production of morphinone from morphine and then nonenzymatic binding of 2-mercaptoethanol to morphinone.  相似文献   

4.
Acetylpyridine NADP replaced NADP in promoting the Mn2+ ion-requiring mitochondrial "malic" enzyme of Hymenolepis diminuta. Disrupted mitochondria displayed low levels of an apparent oxaloacetate-forming malate dehydrogenase activity when NAD or acetylpyridine NAD served as the coenzyme. Significant malate-dependent reduction of acetylpyridine NAD by H. diminuta mitochondria required Mn2+ ion and NADP, thereby indicating the tandem operation of "malic" enzyme and NADPH:NAD transhydrogenase. Incubation of mitochondrial preparations with oxaloacetate resulted in a non-enzymatic decarboxylation reaction. Coupling of malate oxidation with electron transport via the "malic" enzyme and transhydrogenase was demonstrated by polarographic assessment of mitochondrial reduced pyridine nucleotide oxidase activity.  相似文献   

5.
A soluble pyridine nucleotide transhydrogenase (UdhA) has been used to increase the productivity and yield of PHB in vivo. By inducing a high level of UdhA, which can transfer reducing equivalents between NAD and NADP, we have increased NADPH availability, resulting in high yield and productivity of PHB in Escherichia coli. Coexpression of the phb operon from Alcaligenes eutrophus H16 and the native udhA from E. coli from high copy plasmids resulted in an increase in PHB yield from 49 to 66% g of PHB per gram of total cell dry weight and an increase in final concentration from 3.52 to 6.42 g/L; the PHB concentration of the udhA carrying strain is almost twice that of the control strain expressing only the phb operon. The results of this study demonstrate the effectiveness of cofactor manipulation and its application as a tool in metabolic engineering.  相似文献   

6.
Morphine 6-dehydrogenase, which catalyzes the dehydrogenation of morphine to morphinone, was purified 815-fold to a homogeneous protein from the soluble fraction of hamster liver with a yield of 15%. The enzyme was a monomeric protein with a molecular weight of 38 kDa and an isoelectric point of 5.6. Although both NAD and NADP served as cofactors, the enzyme activity with NADP was less than 5% that found with NAD at pH 7.4. With NAD, the enzyme gave the maximal activity at pH 9.3, and the K(m) and V(max) values toward morphine were 1.0 mM and 0.43 unit/mg protein, respectively. Among morphine congeners, normorphine exhibited higher activity than morphine, but codeine and ethylmorphine were poor substrates, and dihydromorphine and dihydrocodeine showed no detectable activity. The enzyme also exhibited significant activity for a variety of cyclic and alicyclic alcohols. In addition to xenobiotics, the enzyme catalyzed the dehydrogenation of 17beta-hydroxysteroids with much higher affinities than morphine. In the reverse reaction, the enzyme exhibited high activity for o-quinones, but morphinone, naloxone, and aromatic aldehydes and ketones were reduced at slow rates. Sulfhydryl reagents and ketamine strongly inhibited the enzyme, whereas pyrazole, barbital, and indomethacin had little effect on enzyme activity. 17beta-Hydroxysteroids inhibited the enzyme in a competitive manner against morphine. A total of 302 amino acid residues, which comprised approximately 94% of whole protein, were identified by sequencing of the peptides obtained by proteolytic digestion. This amino acid sequence of the enzyme showed significant homology to members of the aldo-keto reductase (AKR) superfamily and shared 63-64% identity with members of the AKR1C subfamily. These findings indicate that the enzyme is a new member of the AKR superfamily that is involved in steroid metabolism as 17beta-hydroxysteroid dehydrogenase as well as xenobiotic metabolism.  相似文献   

7.
1. The NAD(P) transhydrogenase activity of the soluble fraction of sonicated rat liver mitochondrial preparations was greater than the NAD-linked isocitrate dehydrogenase activity, and the NAD-linked and NADP-linked isocitrate dehydrogenase activities were not additive. The NAD-linked isocitrate dehydrogenase activity was destroyed by an endogenous autolytic system or by added nucleotide pyrophosphatase, and was restored by a catalytic amount of NADP. 2. We concluded that the isocitrate dehydrogenase of rat liver mitochondria was exclusively NADP-specific, and that the oxoglutarate/isocitrate couple could therefore be used unequivocally as redox reactant for NADP in experiments designed to operate only the NAD(P) transhydrogenase (or loop 0) segment of the respiratory chain in intact mitochondria. 3. During oxidation of isocitrate by acetoacetate in intact, anaerobic, mitochondria via the rhein-sensitive, but rotenone- and arsenite-insensitive, NAD(P) transhydrogenase, measurements of the rates of carbonyl cyanide p-trifluoromethoxyphenylhydrazone-sensitive and carbonyl cyanide p-trifluoromethoxyphenylhydrazone-insensitive pH change in the presence of various oxoglutarate/isocitrate concentration ratios gave an -->H(+)/2e(-) quotient of 1.94+/-0.12 for outward proton translocation by the NAD(P) transhydrogenase. 4. Measurements with a K(+)-sensitive electrode confirmed that the electrogenicity of the NAD(P) transhydrogenase reaction corresponded to the translocation of one positive charge per acid equivalent. 5. Sluggish reversal of the NAD(P) transhydrogenase reaction resulted in a significant inward proton translocation. 6. The possibility that isocitrate might normally be oxidized via loop 0 at a redox potential of -450mV, or even more negative, is discussed, and implies that a P/O quotient of 4 for isocitrate oxidation might be expected.  相似文献   

8.
An enzymatic system for the regeneration of redox cofactors NADH and NADPH was investigated in nanostructural reverse micelles using bacterial glycerol dehydrogenase (GLD) and soluble transhydrogenase (STH). Catalytic conversion of NAD+ to NADH was realized in the sodium dioctylsulfosuccinate (AOT)/isooctane reverse micellar system harboring GLD and a sacrificial substrate, glycerol. The initial rate of NADH regeneration was enhanced by exogenous addition of ammonium sulfate into the reverse micelles, suggesting that NH4+ acts as a monovalent cationic activator. STH was successfully entrapped in the AOT/isooctane reverse micelles as well as GLD and was revealed to be capable of catalyzing the stoichiometric hydrogen transfer reaction between NADP+ and NADPH in reverse micelles. These results indicate that GLD and STH have potential for use in redox cofactor recycling in reverse micelles, which allows the use of catalytic quantities of NAD(P)H in organic media.  相似文献   

9.
The morphine alkaloid hydromorphone (dihydromorphinone) was identified as an intermediary metabolite in the degradation of morphine by Pseudomonas putida M10. A constitutive NADH-dependent morphinone reductase capable of catalyzing the reduction of the 7,8-unsaturated bond of morphinone and codeinone, yielding hydromorphone and hydrocodone, respectively, was shown to be present in cell extracts. The structures have been identified by 1H nuclear magnetic resonance and mass spectrometry. Morphinone reductase has been partially purified by anion-exchange and gel filtration chromatography. This enzyme has potential applications as a biocatalyst for the synthesis of the highly potent analgesic hydromorphone and the antitussive hydrocodone.  相似文献   

10.
Total pyridine nucleotide concentration of root tissue for young soybean (Glycine max var. Bansei) and sunflower (Helianthus annuus L. var. Mammoth Russian) plants is the same with either ammonium or nitrate, but nitrate results in an increased proportion of total oxidized plus reduced NADP (NADP[H]) seemingly at the expense of NAD. The activity of NADH- and NADPH-dependent forms of glutamic acid dehydrogenase is correlated with the ratio of total oxidized plus reduced NAD to NADP(H). The low NAD: NADH ratio maintained in nitrate roots despite active NADH utilization via nitrate reductase and glutamic acid dehydrogenase may be the result of nitrate-stimulated glycolysis. Nitrate roots also maintain a high level of NADPH, presumably by the stimulatory effect of nitrate utilization on glucose-6-phosphate dehydrogenase activity. In the presence of nitrate rather than ammonium, the highly active nitrate-reducing leaves of soybean show a greater proportion of total pyridine nucleotide in the form of NADP(H) than do the inactive leaves of sunflower.  相似文献   

11.
Previous studies have shown that the interaction of P450 reductase with bound NADP(H) is essential to ensure fast electron transfer through the two flavin cofactors. In this study we investigated in detail the interaction of the house fly flavoprotein with NADP(H) and a number of nucleotide analogues. 1,4,5,6-Tetrahydro-NADP, an analogue of NADPH, was used to characterize the interaction of P450 reductase with the reduced nucleotide. This analogue is inactive as electron donor, but its binding affinity and rate constant of release are very close to those for NADPH. The 2'-phosphate contributes about 5 kcal/mol of the binding energy of NADP(H). Oxidized nicotinamide does not interact with the oxidized flavoprotein, while reduced nicotinamide contributes 1.3 kcal/mol of the binding energy. Oxidized P450 reductase binds NADPH with a K(d) of 0.3 microM, while the affinity of the reduced enzyme is considerably lower, K(d) = 1.9 microM. P450 reductase catalyzes a transhydrogenase reaction between NADPH and oxidized nucleotides, such as thionicotinamide-NADP(+), acetylpyridine-NADP(+), or [(3)H]NADP(+). The reverse reaction, reduction of [(3)H]NADP(+) by the reduced analogues, is also catalyzed by P450 reductase. We define the mechanism of the transhydrogenase reaction as follows: NADPH binding, hydride ion transfer, and release of the NADP(+) formed. An NADP(+) or its analogue binds to the two-electron-reduced flavoprotein, and the electron-transfer steps reverse to transfer hydride ion to the oxidized nucleotide, which is released. Measurements of the flavin semiquinone content, rate constant for NADPH release, and transhydrogenase turnover rates allowed us to estimate the steady-state distribution of P450 reductase species during catalysis, and to calculate equilibrium constants for the interconversion of catalytic intermediates. Our results demonstrate that equilibrium redox potentials of the flavin cofactors are not the sole factor governing rapid electron transfer during catalysis, but conformational changes must be considered to understand P450 reductase catalysis.  相似文献   

12.
delta1-Pyrroline-5-carboxylate (PCA) reductase [L-proline:NAD(P)+5-oxidoreductase, EC 1.5.1.2] has been purified over 200-fold from Escherichia coli K-12. It has a molecular weight of approximately 320,000. PCA reductase mediates the pyridine nucleotide-linked reduction of PCA to proline but not the reverse reaction (even at high substrate concentrations). The partially purified preparation is free of competing pyridine nucleotide oxidase, PCA dehydrogenase, and proline oxidase activities. The Michaelis constant (Km) values for the substrate, PCA, with reduced nicotinamide adenine dinucleotide phosphate (NADPH) or NADH as cofactor are 0.15 and 0.14 mM, respectively. The Km values determined for NADPH and NADH are 0.03 and 0.23 mM, respectively. Although either NADPH or NADH can function as cofactor, the activity observed with NADPH is severalfold greater. PCA reductase is not repressed by growth in the presence of proline, but it is inhibited by the reaction end products, proline and NADP.  相似文献   

13.
Glutamate dehydrogenase (GDH) from vertebrates is unusual among NAD(P)H-dependent dehydrogenases in that it can use either NAD(H) or NADP(H) as cofactor. In this study, we measure the rate of cofactor utilization by bovine GDH when both cofactors are present. Methods for both reaction directions were developed, and for the first time, to our knowledge, the GDH activity has been simultaneously studied in the presence of both NAD(H) and NADP(H). Our data indicate that NADP(H) has inhibitory effects on the rate of NAD(H) utilization by GDH, a characteristic of GDH not previously recognized. The response of GDH to allosteric activators in the presence of NAD(H) and NADP(H) suggests that ADP and leucine moderate much of the inhibitory effect of NADP(H) on the utilization of NAD(H). These results illustrate that simple assumptions of cofactor preference by mammalian GDH are incomplete without an appreciation of allosteric effects when both cofactors are simultaneously present.  相似文献   

14.
Nitric oxide reductase (Nor) cytochrome P450nor (P450nor) is unique because it is catalytically self-sufficient, receiving electrons directly from NADH or NADPH. However, little is known about the direct binding of NADH to cytochrome. Here, we report that oxidized pyridine nucleotides (NAD(+) and NADP(+)) and an analogue induce a spectral perturbation in bound heme when mixed with P450nor. The P450nor isoforms are classified according to electron donor specificity for NADH or NADPH. One type (Fnor, a P450nor of Fusarium oxysporum) utilizes only NADH. We found that NAD(+) induced a type I spectral change in Fnor, whereas NADP(+) induced a reverse type I spectral change, although the K(d) values for both were comparable. In contrast, NADP(+) as well as NAD(+) caused a type I spectral change in Tnor, a P450nor isozyme from Trichosporon cutaneum that utilizes both NADH and NADPH as electron donors. The B' helix region of Tnor ((73)SAGGKAAA(80)) contains some Ala and Gly residues, whereas the sequence is replaced at a few sites with more bulky amino acid residues in Fnor ((73)SASGKQAA(80)). A single mutation (S75G) significantly improved the NADPH- dependent Nor activity of Fnor, and the overall activity was accelerated via the NADPH-enhanced reduction step. These results showed that pyridine nucleotide cofactors can bind P450nor and that only a few residues in the B' helix region determine cofactor specificity. We further showed that a poor electron donor (NADPH) could also bind Fnor, but an appropriate configuration for electron transfer is blocked by steric hindrance mainly by Ser(75) against the 2'-phosphate moiety. The present results along with previous observations together revealed a novel motif for cofactor binding.  相似文献   

15.
Akira Kusai  Tateo Yamanaka 《BBA》1973,292(3):621-633
A highly purified preparation of an NAD(P) reductase was obtained from Chlorobium thiosulfatophilum and some of its properties were studied. The enzyme possesses FAD as the prosthetic group, and reduces benzyl viologen, 2,6-dichloro-phenolindophenol and cytochromes c, including cytochrome c-555 (C. thiosulfato-philum), with NADPH or NADH as the electron donor. It reduces NADP+ or NAD+ photosynthetically with spinach chloroplasts in the presence of added spinach ferredoxin. It reduces the pyridine nucleotides with reduced benzyl viologen. The enzyme also shows a pyridine nucleotide transhydrogenase activity. In these reactions, the type of pyridine nucleotide (NADP or NAD) which functions more efficiently with the enzyme varies with the concentration of the nucleotide used; at concentrations lower than approx. 1.0 mM, NADPH (or NADP+) is better electron donor (or acceptor), while NADH (or NAD+) is a better electron donor (or acceptor) at concentrations higher than approx. 1.0 mM. Reduction of dyes or cytochromes c catalysed by the enzyme is strongly inhibited by NADP+, 2′-AMP and and atebrin.  相似文献   

16.
Woodyer R  van der Donk WA  Zhao H 《Biochemistry》2003,42(40):11604-11614
Homology modeling was used to identify two particular residues, Glu175 and Ala176, in Pseudomonas stutzeri phosphite dehydrogenase (PTDH) as the principal determinants of nicotinamide cofactor (NAD(+) and NADP(+)) specificity. Replacement of these two residues by site-directed mutagenesis with Ala175 and Arg176 both separately and in combination resulted in PTDH mutants with relaxed cofactor specificity. All three mutants exhibited significantly better catalytic efficiency for both cofactors, with the best kinetic parameters displayed by the double mutant, which had a 3.6-fold higher catalytic efficiency for NAD(+) and a 1000-fold higher efficiency for NADP(+). The cofactor specificity was changed from 100-fold in favor of NAD(+) for the wild-type enzyme to 3-fold in favor of NADP(+) for the double mutant. Isoelectric focusing of the proteins in a nondenaturing gel showed that the replacement with more basic residues indeed changed the effective pI of the protein. HPLC analysis of the enzymatic products of the double mutant verified that the reaction proceeded to completion using either substrate and produced only the corresponding reduced cofactor and phosphate. Thermal inactivation studies showed that the double mutant was protected from thermal inactivation by both cofactors, while the wild-type enzyme was protected by only NAD(+). The combined results provide clear evidence that Glu175 and Ala176 are both critical for nicotinamide cofactor specificity. The rationally designed double mutant might be useful for the development of an efficient in vitro NAD(P)H regeneration system for reductive biocatalysis.  相似文献   

17.
Circular dichroism has been used to monitor the binding of pyridine nucleotide cofactors to enzyme-folate analog complexes of dihydrofolate reductase from Escherichia coli B (MB 1428). The enzyme binds one molar equivalent of many folate analogs and two molar equivalents of several pyridine nucleotide cofactors. The apo-enzyme has very low optical activity. The binding of folate analogs including folate, dihydrofolate, methotrexate, trimethoprim and pyrimethamine induce large Cotton effects. Pyridine nucleotides when bound to the enzyme-folate analog complexes also induce new optically active bands; all the effects being due to the first molar equivalent of cofactor bound. NADPH and NADP+ induce very similar bands when bound to the enzyme-methotrexate complex suggesting that the geometry of the complexes formed are very similar. The oxidized and reduced cofactor likewise have similar effects on the enzyme-folate complex. However, NADPH and NADP+ addition to both the enzyme-trimethoprim and enzyme-pyrimethamine complexes have significantly different effects on the circular dichroism spectra, suggesting that the inhibitors which are less homologous to the natural dihydrofolate substrate allow more conformational freedom in the enzyme-inhibitor-cofactor complex. In most cases the prior binding of the folate analog greatly increases the binding of the first molar equivalent of cofactor so that at concentrations of approx. 5-20 muM the binding appears stoichiometric. Pyrimethamine is an exception in that it apparently has no effect on the binding of NADPH to the enzyme.  相似文献   

18.
Cofactor specificity in the aldo‐keto reductase (AKR) superfamily has been well studied, and several groups have reported the rational alteration of cofactor specificity in these enzymes. Although most efforts have focused on mesostable AKRs, several putative AKRs have recently been identified from hyperthermophiles. The few that have been characterized exhibit a strong preference for NAD(H) as a cofactor, in contrast to the NADP(H) preference of the mesophilic AKRs. Using the design rules elucidated from mesostable AKRs, we introduced two site‐directed mutations in the cofactor binding pocket to investigate cofactor specificity in a thermostable AKR, AdhD, which is an alcohol dehydrogenase from Pyrococcus furiosus. The resulting double mutant exhibited significantly improved activity and broadened cofactor specificity as compared to the wild‐type. Results of previous pre‐steady‐state kinetic experiments suggest that the high affinity of the mesostable AKRs for NADP(H) stems from a conformational change upon cofactor binding which is mediated by interactions between a canonical arginine and the 2′‐phosphate of the cofactor. Pre‐steady‐state kinetics with AdhD and the new mutants show a rich conformational behavior that is independent of the canonical arginine or the 2′‐phosphate. Additionally, experiments with the highly active double mutant using NADPH as a cofactor demonstrate an unprecedented transient behavior where the binding mechanism appears to be dependent on cofactor concentration. These results suggest that the structural features involved in cofactor specificity in the AKRs are conserved within the superfamily, but the dynamic interactions of the enzyme with cofactors are unexpectedly complex. Biotechnol. Bioeng. 2010;107: 763–774. © 2010 Wiley Periodicals, Inc.  相似文献   

19.
We have previously reported that cytochrome P-450LTB in the microsomes of human polymorphonuclear leukocytes (PMN) catalyzes three omega-oxidations of leukotriene B4 (LTB4), leading to the sequential formation of 20-OH-LTB4, 20-CHO-LTB4, and 20-COOH-LTB4 (Soberman, R.J., Sutyak, J.P., Okita, R.T., Wendelborn, D.F., Roberts, L.J., II, and Austen, K. F. (1988) J. Biol. Chem. 263, 7996-8002). The identification of the novel final intermediate, 20-CHO-LTB4, allowed direct analysis of its metabolism by PMN microsomes in the presence of adenine nucleotide cofactors. Microsomes in the presence of 100 microM NAD+ or 100 microM NADP+ converted 1.0 microM 20-CHO-LTB4 to 20-COOH-LTB4 with a Km of 2.4 +/- 0.8 microM (mean +/- S.E., n = 4) and a Vmax of 813.9 +/- 136.6 pmol.min-1.mg-1, for NAD+, as compared to 0.12 microM and 5.0 pmol.min-1.mg-1 (n = 2) for NADPH as a cofactor. The conversion of 1.0 microM of 20-CHO-LTB4 to 20-COOH-LTB4 in the presence of saturating concentrations (1.0 mM) of both NAD+ and NADP+ was not greater than the reaction in the presence of 1.0 mM of each cofactor separately, indicating that NAD+ and NADP+ were cofactors for the same enzyme. Antibody to cytochrome P-450 reductase did not inhibit the conversion of 20-CHO-LTB4 to 20-COOH-LTB4. When 1.0 microM 20-OH-LTB4 was added to microsomes in the presence of NADPH, approximately three-fourths of the product formed (63.7 +/- 5.1 pmol; mean +/- S.E., n = 3) was 20-CHO-LTB4 and approximately one-fourth (21.3 +/- 3.9 pmol; mean +/- S.E., n = 3) was 20-COOH-LTB4. In the presence of both NADPH and NAD+, only 20-COOH-LTB4 (85.5 +/- 9.9 pmol; mean +/- S.E., n = 3) was formed. PMN microsomes also contain an NADH-dependent aldehyde reductase which converts 20-CHO-LTB4 to 20-OH-LTB4, a member of the LTB4 family of molecules with biological activity. Based upon kinetic, cofactor and inhibition data, microsomal aldehyde dehydrogenase preferentially regulates the final and irreversible inactivation step in the LTB4 metabolic sequence.  相似文献   

20.
Abstract Cell-free extracts of the photosynthetic eubacterium Rhodomicrobium vannielii contained both NADP and NAD-linked isocitrate dehydrogenase activities. Apparent K m values of 12 μM for NADP, 0.75 mM for NAD, 9.3 μM for isocitrate (NADP utilising) and 8.2 μM for isocitrate (NAD utilising) were determined in such extracts. Four lines of evidence indicated that one enzyme was responsible for the two activities; (i) non-additivity of reaction rates in the presence of both NADP and NAD (ii) the presence of one band which stained for activity with both cofactors on non-denaturing polyacrylamide gels (iii) identical heat inactivation kinetics for both activities (iv) co-elution of both activities after ion-exchange and hydrophobic interaction chromatography. This is the first report of a eubacterial isocitrate dehydrogenase with dual cofactor specificity.  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