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.
Purification and characterization of enzymes metabolizing retinaldehyde, propionaldehyde, and octanaldehyde from four human livers and three kidneys were done to identify enzymes metabolizing retinaldehyde and their relationship to enzymes metabolizing other aldehydes. The tissue fractionation patterns from human liver and kidney were the same, indicating presence of the same enzymes in human liver and kidney. Moreover, in both organs the major NAD(+)-dependent retinaldehyde activity copurified with the propionaldehyde and octanaldehyde activities; in both organs the major NAD(+)-dependent retinaldehyde activity was associated with the E1 isozyme (coded for by aldh1 gene) of human aldehyde dehydrogenase. A small amount of NAD(+)-dependent retinaldehyde activity was associated with the E2 isozyme (product of aldh2 gene) of aldehyde dehydrogenase. Some NAD(+)-independent retinaldehyde activity in both organs was associated with aldehyde oxidase, which could be easily separated from dehydrogenases. Employing cellular retinoid-binding protein (CRBP), purified from human liver, demonstrated that E1 isozyme (but not E2 isozyme) could utilize CRBP-bound retinaldehyde as substrate, a feature thought to be specific to retinaldehyde dehydrogenases. This is the first report of CRBP-bound retinaldehyde functioning as substrate for aldehyde dehydrogenase of broad substrate specificity. Thus, it is concluded that in the human organism, retinaldehyde dehydrogenase (coded for by raldH1 gene) and broad substrate specificity E1 (a member of EC 1. 2.1.3 aldehyde dehydrogenase family) are the same enzyme. These results suggest that the E1 isozyme may be more important to alcoholism than the acetaldehyde-metabolizing enzyme, E2, because competition between acetaldehyde and retinaldehyde could result in abnormalities associated with vitamin A metabolism and alcoholism.  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.
Yeast (Saccharomyces cerevisiae) alcohol dehydrogenase I (SceADH) binds NAD+ and NADH less tightly and turns over substrates more rapidly than does horse (Equus caballus) liver alcohol dehydrogenase E isoenzyme (EcaADH), and neither enzyme uses NADP efficiently. Amino acid residues in the proposed adenylate binding pocket of SceADH were substituted in attempts to improve affinity for coenzymes or reactivity with NADP. Substitutions in SceADH (Gly202Ile or Ser246Ile) with the corresponding residues in the adenine binding site of the homologous EcaADH have modest effects on coenzyme binding and other kinetic constants, but the Ser246Ile substitution decreases turnover numbers by 350-fold. The Ser176Phe substitution (also near adenine site) significantly decreases affinity for coenzymes and turnover numbers. In the consensus nucleotide-binding betaalphabeta fold sequence, SceADH has two alanine residues (177-GAAGGLG-183) instead of the Leu200 in EcaADH (199-GLGGVG-204); the Ala178-Ala179 to Leu substitution significantly decreases affinity for coenzymes and turnover numbers. Some NADP-dependent enzymes have an Ala corresponding to Gly183 in SceADH; the Gly183Ala substitution significantly decreases affinity for coenzymes and turnover numbers. NADP-dependent enzymes usually have a neutral residue instead of the Asp (Asp201 in SceADH) that interacts with the hydroxyl groups of the adenosine ribose, along with a basic residue (at position 202 or 203) to stabilize the 2'-phosphate of NADP. The Gly203Arg change in SceADH does not significantly affect the kinetics. The Gly183Ala or Gly203Arg substitutions do not enable SceADH to use NADP+ as coenzyme. SceADH with the single Asp201Gly or double Asp201Gly:Gly203Arg substitutions have similar, low activity with NADP+. The results suggest that several of the amino acid residues participate in coenzyme binding and that conversion of specificity for coenzyme requires multiple substitutions.  相似文献   

4.
Sharkey MA  Gori A  Capone M  Engel PC 《The FEBS journal》2012,279(17):3003-3009
Active-site mutants of glutamate dehydrogenase from Clostridium?symbiosum have been designed and constructed and the effects on coenzyme preference evaluated by detailed kinetic measurements. The triple mutant F238S/P262S/D263K shows complete reversal in coenzyme selectivity from NAD(H) to NADP(H) with retention of high levels of catalytic activity for the new coenzyme. For oxidized coenzymes, k(cat) /K(m) ratios of the wild-type and triple mutant enzyme indicate a shift in preference of approximately 1.6?×?10(7) -fold, from ~?80?000-fold in favour of NAD(+) to ~?200-fold in favour of NADP(+) . For reduced coenzymes the corresponding figure is 1.7?×?10(4) -fold, from ~?1000-fold in favour of NADH to ~?17-fold in favour of NADPH. A fourth mutation (N290G), previously identified as having a potential bearing on coenzyme specificity, did not engender any further shift in preference when incorporated into the triple mutant, despite having a significant effect when expressed as a single mutant.  相似文献   

5.
The C-terminal domain (C(t)-FDH) of 10-formyltetrahydrofolate dehydrogenase (FDH, ALDH1L1) is an NADP(+)-dependent oxidoreductase and a structural and functional homolog of aldehyde dehydrogenases. Here we report the crystal structures of several C(t)-FDH mutants in which two essential catalytic residues adjacent to the nicotinamide ring of bound NADP(+), Cys-707 and Glu-673, were replaced separately or simultaneously. The replacement of the glutamate with an alanine causes irreversible binding of the coenzyme without any noticeable conformational changes in the vicinity of the nicotinamide ring. Additional replacement of cysteine 707 with an alanine (E673A/C707A double mutant) did not affect this irreversible binding indicating that the lack of the glutamate is solely responsible for the enhanced interaction between the enzyme and the coenzyme. The substitution of the cysteine with an alanine did not affect binding of NADP(+) but resulted in the enzyme lacking the ability to differentiate between the oxidized and reduced coenzyme: unlike the wild-type C(t)-FDH/NADPH complex, in the C707A mutant the position of NADPH is identical to the position of NADP(+) with the nicotinamide ring well ordered within the catalytic center. Thus, whereas the glutamate restricts the affinity for the coenzyme, the cysteine is the sensor of the coenzyme redox state. These conclusions were confirmed by coenzyme binding experiments. Our study further suggests that the binding of the coenzyme is additionally controlled by a long-range communication between the catalytic center and the coenzyme-binding domain and points toward an α-helix involved in the adenine moiety binding as a participant of this communication.  相似文献   

6.
The beta-subunit of the voltage-sensitive K(+) channels shares 15-30% amino acid identity with the sequences of aldo-keto reductases (AKR) genes. However, the AKR properties of the protein remain unknown. To begin to understand its oxidoreductase properties, we examine the pyridine coenzyme binding activity of the protein in vitro. The cDNA of K(v)beta2.1 from rat brain was subcloned into a prokaryotic expression vector and overexpressed in Escherichia coli. The purified protein was tetrameric in solution as determined by size exclusion chromatography. The protein displayed high affinity binding to NADPH as determined by fluorometric titration. The K(D) values for NADPH of the full-length wild-type protein and the N-terminus deleted protein were 0.1+/-0.007 and 0.05+/-0.006 M, respectively - indicating that the cofactor binding domain is restricted to the C-terminus, and is not drastically affected by the absence of the N-terminus amino acids, which form the ball and chain regulating voltage-dependent inactivation of the alpha-subunit. The protein displayed poor affinity for other coenzymes and the corresponding values of the K(D) for NADH and NAD were between 1-3 microM whereas the K(D) for FAD was >10 microM. However, relatively high affinity binding was observed with 3-acetyl pyridine NADP, indicating selective recognition of the 2' phosphate at the binding site. The selectivity of K(v)beta2.1 for NADPH over NADP may be significant in regulating the K(+) channels as a function of the cellular redox state.  相似文献   

7.
L Zhang  B Ahvazi  R Szittner  A Vrielink  E Meighen 《Biochemistry》1999,38(35):11440-11447
The fatty aldehyde dehydrogenase from the luminescent bacterium, Vibrio harveyi (Vh-ALDH), is unique with respect to its high specificity for NADP(+) over NAD(+). By mutation of a single threonine residue (Thr175) immediately downstream of the beta(B) strand in the Rossmann fold, the nucleotide specificity of Vh-ALDH has been changed from NADP(+) to NAD(+). Replacement of Thr175 by a negatively charged residue (Asp or Glu) resulted in an increase in k(cat)/K(m) for NAD(+) relative to that for NADP(+) of up to 5000-fold due to a decrease for NAD(+) and an increase for NADP(+) in their respective Michaelis constants (K(a)). Differential protection by NAD(+) and NADP(+) against thermal inactivation and comparison of the dissociation constants of NMN, 2'-AMP, 2'5'-ADP, and 5'-AMP for these mutants and the wild-type enzyme clearly support the change in nucleotide specificity. Moreover, replacement of Thr175 with polar residues (N, S, or Q) demonstrated that a more efficient NAD(+)-dependent enzyme T175Q could be created without loss of NADP(+)-dependent activity. Analysis of the three-dimensional structure of Vh-ALDH with bound NADP(+) showed that the hydroxyl group of Thr175 forms a hydrogen bond to the 2'-phosphate of NADP(+). Replacement with glutamic acid or glutamine strengthened interactions with NAD(+) and indicated why threonine would be the preferred polar residue at the nucleotide recognition site in NADP(+)-specific aldehyde dehydrogenases. These results have shown that the size and the structure of the residue at the nucleotide recognition site play the key roles in differentiating between NAD(+) and NADP(+) interactions while the presence of a negative charge is responsible for the decrease in interactions with NADP(+) in Vh-ALDH.  相似文献   

8.
Ferredoxin-NADP(+) reductase catalyses NADP(+) reduction, being specific for NADP(+)/H. To understand coenzyme specificity determinants and coenzyme specificity reversion, mutations at the NADP(+)/H pyrophosphate binding and of the C-terminal regions have been simultaneously introduced in Anabaena FNR. The T155G/A160T/L263P/Y303S mutant was produced. The mutated enzyme presents similar k(cat) values for NADPH and NADH, around 2.5 times slower than that reported for WT FNR with NADPH. Its K(m) value for NADH decreased 20-fold with regard to WT FNR, whereas the K(m) for NADPH remains similar. The combined effect is a much higher catalytic efficiency for NAD(+)/H, with a minor decrease of that for NADP(+)/H. In the mutated enzyme, the specificity for NADPH versus NADH has been decreased from 67,500 times to only 12 times, being unable to discriminate between both coenzymes. Additionally, giving the role stated for the C-terminal Tyr in FNR, its role in the energetics of the FAD binding has been analysed.  相似文献   

9.
The widespread, long sought-after bacterial aerobic phenylalanine/phenylacetate catabolic pathway has recently been elucidated. It proceeds via coenzyme A (CoA) thioesters and involves the epoxidation of the aromatic ring of phenylacetyl-CoA, subsequent isomerization to an uncommon seven-membered C-O-heterocycle (oxepin-CoA), and non-oxygenolytic ring cleavage. Here we characterize the hydrolytic oxepin-CoA ring cleavage catalyzed by the bifunctional fusion protein PaaZ. The enzyme consists of a C-terminal (R)-specific enoyl-CoA hydratase domain (formerly MaoC) that cleaves the ring and produces a highly reactive aldehyde and an N-terminal NADP(+)-dependent aldehyde dehydrogenase domain that oxidizes the aldehyde to 3-oxo-5,6-dehydrosuberyl-CoA. In many phenylacetate-utilizing bacteria, the genes for the pathway exist in a cluster that contains an NAD(+)-dependent aldehyde dehydrogenase in place of PaaZ, whereas the aldehyde-producing hydratase is encoded outside of the cluster. If not oxidized immediately, the reactive aldehyde condenses intramolecularly to a stable cyclic derivative that is largely prevented by PaaZ fusion in vivo. Interestingly, the derivative likely serves as the starting material for the synthesis of antibiotics (e.g. tropodithietic acid) and other tropone/tropolone related compounds as well as for ω-cycloheptyl fatty acids. Apparently, bacteria made a virtue out of the necessity of disposing the dead-end product with ring hydrolysis as a metabolic branching point.  相似文献   

10.
11.
1. The inducibility of hepatic cytosolic aldehyde dehydrogenase activity was studied in rat, mouse, guinea pig, chicken, frog, salamander and rainbow trout, by using two different types of inducers of drug metabolism. 2. Phenobarbital (a type I inducer of drug metabolizing enzymes) increased total liver cytosolic aldehyde dehydrogenase activity (up to 20-fold) in a genetically defined substrain of responsive rats (RR) and only slightly, if at all, in a non-responsive substrain (rr). On the contrary, both types of rats showed a highly induced aldehyde dehydrogenase activity after treatment with methylcholanthrene (a type II inducer). Phenobarbital is affecting mainly an isozyme of aldehyde dehydrogenase which is best measured with propionaldehyde as the substrate and NAD as the coenzyme (P/NAD). 3. Administration of phenobarbital to mice produced only a slight increase (2-fold) in the P/NAD aldehyde dehydrogenase activity. 4. Methylcholanthrene treatment caused a 2-fold increase of the hepatic P/NAD aldehyde dehydrogenase activity in the chicken. 5. In the guinea pig, phenobarbital produced an approximate 3-fold increase of the P/NAD activity. Methylcholanthrene had a similar effect, although to a lesser extent. 6. In the salamander, a 4-fold increase was detected in the enzyme activity measured with benzaldehyde as the substrate and NADP as the coenzyme (B/NADP), after treatment with either phenobarbital or methylcholanthrene. 7. The hepatic aldehyde dehydrogenase activities were found unchanged in the rainbow trout, after treatment with phenobarbital or 2,3,7,8-tetrachlorodibenzo-p-dioxin. 8. The rat model remains the only one examined that shares with human hepatocytes strong inducibility of the B/NADP aldehyde dehydrogenase isozyme upon treatment with polycyclic aromatic hydrocarbons.  相似文献   

12.
Among pyridine-nucleotide-dependent oxidoreductases, the class 3 family of aldehyde dehydrogenases (ALDHs) is unusual in its ability to function with either NAD or NADP. This is all the more surprising because an acidic residue, Glu140, coordinates the adenine ribose 2' hydroxyl. In many NAD-dependent dehydrogenases a similarly placed carboxylate is thought to be responsible for exclusion of NADP. The corresponding residue in most (approximately 71%) sequences in the ALDH extended family is also Glu, and most of these are NAD-specific enzymes. Site-directed mutagenesis was performed on this residue in rat class 3 ALDH. Our results indicate that this residue contributes to tighter binding of NAD in the native enzyme, but suggest that additional factors must contribute to the ability to utilize NADP. Mutagenesis of an adjacent basic residue (Lys137) indicates that it is even more essential for binding both coenzymes, consistent with its conservation in nearly all ALDHs (> 98%).  相似文献   

13.
1. Kinetic studies of glutamate dehydrogenase were made with wide concentration ranges of the coenzymes NAD(+) and NADP(+) and the substrates glutamate and norvaline. Initial-rate parameters were evaluated. 2. Deviations from Michaelis-Menten behaviour towards higher activity were observed with increasing concentrations of either coenzyme with glutamate as substrate, but not with norvaline as substrate. 3. In phosphate buffer, pH7.0, Lineweaver-Burk plots with either coenzyme as variable and a constant, large glutamate concentration showed three or four linear regions of different slope with relatively sharp discontinuities. Maximum rates obtained by extrapolation and Michaelis constants for the coenzymes increased in steps with increase of coenzyme concentration. 4. In the absence of evidence of heterogeneity of the enzyme and coenzyme preparations, the results are interpreted in terms of negative homotropic interactions between the enzyme subunits. It is suggested that sharp discontinuities in Lineweaver-Burk plots or reciprocal binding plots may be characteristic of this new type of interaction, which can be explained in terms of an Adair-Koshland model, but not by the model of Monod, Wyman & Changeux.  相似文献   

14.
6-Phosphogluconate dehydrogenase is a potential target for new drugs against African trypanosomiasis. Phosphorylated aldonic acids are strong inhibitors of 6-phosphogluconate dehydrogenase, and 4-phospho-d-erythronate (4PE) and 4-phospho-d-erythronohydroxamate are two of the strongest inhibitors of the Trypanosoma brucei enzyme. Binding of the substrate 6-phospho-d-gluconate (6PG), the inhibitors 5-phospho-d-ribonate (5PR) and 4PE, and the coenzymes NADP, NADPH and NADP analogue 3-amino-pyridine adenine dinucleotide phosphate to 6-phospho-d-gluconate dehydrogenase from T. brucei was studied using isothermal titration calorimetry. Binding of the substrate (K(d) = 5 microm) and its analogues (K(d) =1.3 microm and K(d) = 2.8 microm for 5PR and 4PE, respectively) is entropy driven, whereas binding of the coenzymes is enthalpy driven. Oxidized coenzyme and its analogue, but not reduced coenzyme, display a half-site reactivity in the ternary complex with the substrate or inhibitors. Binding of 6PG and 5PR poorly affects the dissociation constant of the coenzymes, whereas binding of 4PE decreases the dissociation constant of the coenzymes by two orders of magnitude. In a similar manner, the K(d) value of 4PE decreases by two orders of magnitude in the presence of the coenzymes. The results suggest that 5PR acts as a substrate analogue, whereas 4PE mimics the transition state of dehydrogenation. The stronger affinity of 4PE is interpreted on the basis of the mechanism of the enzyme, suggesting that the inhibitor forces the catalytic lysine 185 into the protonated state.  相似文献   

15.
Nidetzky B  Klimacek M  Mayr P 《Biochemistry》2001,40(34):10371-10381
Microbial xylose reductase, a representative aldo-keto reductase of primary sugar metabolism, catalyzes the NAD(P)H-dependent reduction of D-xylose with a turnover number approximately 100 times that of human aldose reductase for the same reaction. To determine the mechanistic basis for that physiologically relevant difference and pinpoint features that are unique to the microbial enzyme among other aldo/keto reductases, we carried out stopped-flow studies with wild-type xylose reductase from the yeast Candida tenuis. Analysis of transient kinetic data for binding of NAD(+) and NADH, and reduction of D-xylose and oxidation of xylitol at pH 7.0 and 25 degrees C provided estimates of rate constants for the following mechanism: E + NADH right arrow over left arrow E.NADH right arrow over left arrow E.NADH + D-xylose right arrow over left arrow E.NADH.D-xylose right arrow over left arrow E.NAD(+).xylitol right arrow over left arrow E.NAD(+) right arrow over left arrow E.NAD(+) right arrow over left arrow E + NAD(+). The net rate constant of dissociation of NAD(+) is approximately 90% rate limiting for k(cat) of D-xylose reduction. It is controlled by the conformational change which precedes nucleotide release and whose rate constant of 40 s(-)(1) is 200 times that of completely rate-limiting E.NADP(+) --> E.NADP(+) step in aldehyde reduction catalyzed by human aldose reductase [Grimshaw, C. E., et al. (1995) Biochemistry 34, 14356-14365]. Hydride transfer from NADH occurs with a rate constant of approximately 170 s(-1). In reverse reaction, the E.NADH --> E.NADH step takes place with a rate constant of 15 s(-1), and the rate constant of ternary-complex interconversion (3.8 s(-1)) largely determines xylitol turnover (0.9 s(-1)). The bound-state equilibrium constant for C. tenuis xylose reductase is estimated to be approximately 45 (=170/3.8), thus greatly favoring aldehyde reduction. Formation of productive complexes, E.NAD(+) and E.NADH, leads to a 7- and 9-fold decrease of dissociation constants of initial binary complexes, respectively, demonstrating that 12-fold differential binding of NADH (K(i) = 16 microM) vs NAD(+) (K(i) = 195 microM) chiefly reflects difference in stabilities of E.NADH and E.NAD(+). Primary deuterium isotope effects on k(cat) and k(cat)/K(xylose) were, respectively, 1.55 +/- 0.09 and 2.09 +/- 0.31 in H(2)O, and 1.26 +/- 0.06 and 1.58 +/- 0.17 in D(2)O. No deuterium solvent isotope effect on k(cat)/K(xylose) was observed. When deuteration of coenzyme selectively slowed the hydride transfer step, (D)()2(O)(k(cat)/K(xylose)) was inverse (0.89 +/- 0.14). The isotope effect data suggest a chemical mechanism of carbonyl reduction by xylose reductase in which transfer of hydride ion is a partially rate-limiting step and precedes the proton-transfer step.  相似文献   

16.
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   

17.
Benzoate, a strategic intermediate in aerobic aromatic metabolism, is metabolized in various bacteria via an unorthodox pathway. The intermediates of this pathway are coenzyme A (CoA) thioesters throughout, and ring cleavage is nonoxygenolytic. The fate of the ring cleavage product 3,4-dehydroadipyl-CoA semialdehyde was studied in the beta-proteobacterium Azoarcus evansii. Cell extracts contained a benzoate-induced, NADP(+)-specific aldehyde dehydrogenase, which oxidized this intermediate. A postulated putative long-chain aldehyde dehydrogenase gene, which might encode this new enzyme, is located on a cluster of genes encoding enzymes and a transport system required for aerobic benzoate oxidation. The gene was expressed in Escherichia coli, and the maltose-binding protein-tagged enzyme was purified and studied. It is a homodimer composed of 54 kDa (without tag) subunits and was confirmed to be the desired 3,4-dehydroadipyl-CoA semialdehyde dehydrogenase. The reaction product was identified by nuclear magnetic resonance spectroscopy as the corresponding acid 3,4-dehydroadipyl-CoA. Hence, the intermediates of aerobic benzoyl-CoA catabolic pathway recognized so far are benzoyl-CoA; 2,3-dihydro-2,3-dihydroxybenzoyl-CoA; 3,4-dehydroadipyl-CoA semialdehyde plus formate; and 3,4-dehydroadipyl-CoA. The further metabolism is thought to lead to 3-oxoadipyl-CoA, the intermediate at which the conventional and the unorthodox pathways merge.  相似文献   

18.
Phenylacetaldehyde dehydrogenase (PAD) and lactaldehyde dehydrogenase (ALD) share some structural and kinetic properties. One difference is that PAD can use NAD+ and NADP+, whereas ALD only uses NAD+. An acidic residue has been involved in the exclusion of NADP+ from the active site in pyridine nucleotide-dependent dehydrogenases. However, other factors may participate in NADP+ exclusion. In the present work, analysis of the sequence of the region involved in coenzyme binding showed that residue F180 of ALD might participate in coenzyme specificity. Interestingly, F180T mutation rendered an enzyme (ALD-F180T) with the ability to use NADP+. This enzyme showed an activity of 0.87 micromol/(min * mg) and K(m) for NADP+ of 78 microM. Furthermore, ALD-F180T exhibited a 16-fold increase in the V(m) /K(m) ratio with NAD+ as the coenzyme, from 12.8 to 211. This increase in catalytic efficiency was due to a diminution in K(m) for NAD+ from 47 to 7 microM and a higher V(m) from 0.51 to 1.48 micromol/(min * mg). In addition, an increased K(d) for NADH from 175 (wild-type) to 460 microM (mutant) indicates a faster product release and possibly a change in the rate-limiting step. For wild-type ALD it is described that the rate-limiting step is shared between deacylation and coenzyme dissociation. In contrast, in the present report the rate-limiting step in ALD-F180T was determined to be exclusively deacylation. In conclusion, residue F180 participates in the exclusion of NADP+ from the coenzyme binding site and disturbs the binding of NAD+.  相似文献   

19.
Aminopropionaldehyde dehydrogenase was purified to apparent homogeneity from 1,3-diaminopropane-grown cells of Arthrobacter sp. TMP-1. The native molecular mass and the subunit molecular mass of the enzyme were approximately 20,5000 and 52,000, respectively, suggesting that the enzyme is a tetramer of identical subunits. The apparent Michaelis constant (K(m)) for 1,3-diaminopropane was approximately 3 microM. The enzyme equally used both NAD(+) and NADP(+) as coenzymes. The apparent K(m) values for NAD(+) and NADP(+) were 255 microM and 108 microM, respectively. The maximum reaction rates (V(max)) for NAD(+) and NADP(+) were 102 and 83.3 micromol min(-1) mg(-1), respectively. Some tested aliphatic aldehydes and aromatic aldehydes were inert as substrates. The optimum pH was 8.0-8.5. The enzyme was sensitive to sulfhydryl group-modifying reagents.  相似文献   

20.
Biosensors for the oxidized substrates of NAD(P)(+)-specific dehydrogenases demand the reductive recycling of the coenzymes. So far, suitable catalysts for the corresponding two-electron transfer are not available. In the present paper, this transport has been realized by a combined electrocatalytical and electroenzymatic process. Lipoic acid has been reduced on graphite electrodes functionalized with Fe(II)-phthalocyanine in 95% yield at-1200 mV in phosphate buffer pH 7.0. With the electrocatalytically reduced product, dihydrolipoic acid, lipoamide dehydrogenase could reduce NAD(+) in 20% yield and thioredoxin reductase NADP(+) in 18.4% yield. So far, the combined electrocatalytic/electroenzymatic system has not yet been realized, mainly because at the potential needed for the lipoic acid reduction, a parallel one-electron reduction of NAD(P)(+) was observed, implying the dimerization of the coenzyme.  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