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 31 毫秒
1.
Clostridial glutamate dehydrogenase mutants, designed to accommodate the 2'-phosphate of disfavoured NADPH, showed the expected large specificity shifts with NAD(P)H. Puzzlingly, similar assays with oxidized cofactors initially revealed little improvement with NADP(+) , although rates with NAD(+) were markedly diminished. This article reveals that the enzyme's discrimination in favour of NAD(+) and against NADP(+) had been greatly underestimated and has indeed been abated by a factor of >?16,000 by the mutagenesis. Initially, stopped-flow studies of the wild-type enzyme showed a burst increase of A(340) with NADP(+) but not NAD(+), with amplitude depending on the concentration of the coenzyme, rather than enzyme. Amplitude also varied with the commercial source of the NADP(+). FPLC, HPLC and mass spectrometry identified NAD(+) contamination ranging from 0.04 to 0.37% in different commercial samples. It is now clear that apparent rates of NADP(+) utilization mainly reflected the reduction of contaminating NAD(+), creating an entirely false view of the initial coenzyme specificity and also of the effects of mutagenesis. Purification of the NADP(+) eliminated the burst. With freshly purified NADP(+), the NAD(+) : NADP(+) activity ratio under standard conditions, previously estimated as 300 : 1, is 11,000. The catalytic efficiency ratio is even higher at 80,000. Retested with pure cofactor, mutants showed marked specificity shifts in the expected direction, for example, 16 200 fold change in catalytic efficiency ratio for the mutant F238S/P262S, confirming that the key structural determinants of specificity have been successfully identified. Of wider significance, these results underline that, without purification, even the best commercial coenzyme preparations are inadequate for such studies.  相似文献   

2.
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   

3.
NAD激酶催化辅酶Ⅰ[NAD(H)]发生磷酸化,转变成辅酶Ⅱ[NADP(H)],而还原态辅酶Ⅱ(NADPH)是L-异亮氨酸合成的必要辅因子。为了提高NADPH的供应,首先克隆了谷氨酸棒杆菌NAD激酶基因ppnK,并利用大肠杆菌-棒状杆菌诱导型穿梭表达载体pDXW-8和组成型穿梭表达载体pDXW-9在L-异亮氨酸合成菌——乳糖发酵短杆菌JHI3-156中进行表达。摇瓶发酵后,ppnK诱导表达菌JHI3-156/pDXW-8-ppnK的NAD激酶酶活(4.33±0.74 U/g)比pDXW-8空载菌提高了83.5%,辅酶Ⅱ与辅酶Ⅰ的比例提高了63.8%,L-异亮氨酸产量(3.86±0.12 g/L)提高了82.9%;ppnK组成表达菌JHI3-156/pDXW-9-ppnK的NAD激酶酶活(7.67±0.65 U/g)比pDXW-9空载菌提高了2.20倍,辅酶Ⅱ与辅酶Ⅰ的比例提高了1.34倍,NADPH含量提高了21.7%,L-异亮氨酸产量(2.99±0.18 g/L)提高了41.7%。这说明NAD激酶有助于辅酶Ⅱ的供应和L-异亮氨酸的生物合成,这对于其他氨基酸的生产也有一定的参考依据。  相似文献   

4.
Asymmetric biosyntheses allow for an efficient production of chiral building blocks. The application of whole cells as biocatalysts for asymmetric syntheses is advantageous because they already contain the essential coenzymes NAD(H) or NADP(H), which additionally can be regenerated in the cells. Unfortunately, reduced catalytic activity compared to the oxidoreductase activity is observed in many cases during whole‐cell biotransformation. This may be caused by low intracellular coenzyme pool sizes and/or a decline in intracellular coenzyme concentrations. To enhance the intracellular coenzyme pool sizes, the effects of the precursor metabolites adenine and nicotinic acid on the intracellular accumulation of NAD(H) and NADP(H) were studied in Saccharomyces cerevisiae. Based on the results of simple batch experiments with different precursor additions, fed‐batch processes for the production of yeast cells with enhanced NAD(H) or enhanced NADP(H) pool sizes were developed. Supplementation of the feed medium with 95 mM adenine and 9.5 mM nicotinic acid resulted in an increase of the intracellular NAD(H) concentration by a factor of 10 at the end of the fed‐batch process compared to the reference process. The final NAD(H) concentration remains unchanged if the feed medium was solely supplemented with 95 mM adenine, but intracellular NADP(H) was increased by a factor of 4. The effects of NADP(H) pool sizes on the asymmetric reduction of ethyl‐4‐chloro acetoacetate (CAAE) to the corresponding (S)‐4‐chloro‐3‐hydroxybutanoate (S‐CHBE) was evaluated with S. cerevisiae FasB His6 as an example. An intracellular threshold concentration above 0.07 mM NADP(H) was sufficient to increase the biocatalytic S‐CHBE productivity by 25 % compared to lower intracellular NADP(H) concentrations.  相似文献   

5.
An isocitrate dehydrogenase able to function with either NADP or NAD as coenzyme was purified to homogeneity from cell-free extracts of the purple photosynthetic eubacterium Rhodomicrobium vannielii using a rapid two-step procedure involving dye-ligand affinity chromatography. The enzyme was obtained in 60% yield with specific activities of 23 U.mg protein-1 (NADP-linked reaction) and 18.5 U.mg protein-1 (NAD-linked reaction). The purified enzyme was monomeric and migrated with an approximate Mr of 75,000-80,000 on both SDS/PAGE and non-denaturing PAGE. Affinity constants (Km values) of 2.5 microM for NADP and 0.77 mM for NAD and values for kcat/Km of 981,200 min-1.mM-1 (NADP) and 2455 min-1.mM-1 (NAD) indicated a greater specificity for NADP compared to NAD. A number of metabolites were examined for possible differential regulatory effects on the NADP- and NAD-linked reactions, using a dual-wavelength assay. Oxaloacetate was found to be an effective inhibitor of both reactions and the enzyme was also sensitive to concerted inhibition by glyoxylate and oxaloacetate. The amino-acid composition and the identity of 39 residues at the N-terminus were determined and compared to other isocitrate dehydrogenases. The results suggested a relationship between the Rm. vannielii enzyme and the monomeric isocitrate dehydrogenase isoenzyme II from Vibrio ABE-1.  相似文献   

6.
F Fan  J A Lorenzen  B V Plapp 《Biochemistry》1991,30(26):6397-6401
In the three-dimensional structures of enzymes that bind NAD or FAD, there is an acidic residue that interacts with the 2'- and 3'-hydroxyl groups of the adenosine ribose of the coenzyme. The size and charge of the carboxylate might repel the binding of the 2'-phosphate group of NADP and explain the specificity for NAD. In the NAD-dependent alcohol dehydrogenases, Asp-223 (horse liver alcohol dehydrogenase sequence) appears to have this role. The homologous residue in yeast alcohol dehydrogenase I (residue 201 in the protein sequence) was substituted with Gly, and the D223G enzyme was expressed in yeast, purified, and characterized. The wild-type enzyme is specific for NAD. In contrast, the D223G enzyme bound and reduced NAD+ and NADP+ equally well, but, relative to wild-type enzyme, the dissociation constant for NAD+ was increased 17-fold, and the reactivity (V/K) on ethanol was decreased to 1%. Even though catalytic efficiency was reduced, yeast expressing the altered or wild-type enzyme grew at comparable rates, suggesting that equilibration of NAD and NADP pools is not lethal. Asp-223 participates in binding NAD and in excluding NADP, but it is not the only residue important for determining specificity for coenzyme.  相似文献   

7.
The NAD- and NADP-dependent aldehyde dehydrogenase (ALDH) activities were evaluated in two rat hepatoma cell lines, namely the well-differentiated MH1C1 line and the less differentiated HTC line. Each activity was determined in parallel in isolated rat hepatocytes, for comparison. The aliphatic aldehyde acetaldehyde (ACA) and the aromatic aldehyde benzaldehyde (BA) were used as substrates. With the first substrate the ALDH activities found in the crude cytoplasmic extracts were lower in hepatoma cells than in normal hepatocytes, especially when measured with NADP as coenzyme (ACA/NADP). Otherwise, with benzaldehyde as substrate the NAD-dependent enzyme activity (BA/NAD) was increased about 9-fold in HTC cells over hepatocytes and decreased in MH1C1 cells, while the NADP-dependent (BA/NADP) activity was increased 38- and 2.5-fold in HTC and MH1C1 cell lines, respectively. Studies on the subcellular distribution of these enzyme activities showed that the activity measured with acetaldehyde and NAD (ACA/NAD) was almost equally distributed between the cytosol and the subcellular particles in the three cell populations, but the ACA/NADP activity was shifted towards the cytosolic compartment in hepatomas, especially in HTC cells. The BA/NAD and BA/NADP ALDH activities found in the organelles of hepatoma cells were markedly reduced in comparison with hepatocytes, in favour of the cytosol. The most striking difference between the normal and the transformed cells was the 94-fold increase over hepatocytes of the BA/NADP activity, found in the cytosolic fractions of HTC cells. MH1C1 cells showed a less pronounced (7.5-fold) enhancement of this tumour-associated specific activity.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

8.
Initial-rate studies were made of the oxidation of L-glutamate by NAD+ and NADP+ catalysed by highly purified preparations of dogfish liver glutamate dehydrogenase. With NAD+ as coenzyme the kinetics show the same features of coenzyme activation as seen with the bovine liver enzyme [Engel & Dalziel (1969) Biochem. J. 115, 621--631]. With NADP+ as coenzyme, initial rates are much slower than with NAD+, and Lineweaver--Burk plots are linear over extended ranges of substrate and coenzyme concentration. Stopped-flow studies with NADP+ as coenzyme give no evidence for the accumulation of significant concentrations of NADPH-containing complexes with the enzyme in the steady state. Protection studies against inactivation by pyridoxal 5'-phosphate indicate that NAD+ and NADP+ give the same degree of protection in the presence of sodium glutarate. The results are used to deduce information about the mechanism of glutamate oxidation by the enzyme. Initial-rate studies of the reductive amination of 2-oxoglutarate by NADH and NADPH catalysed by dogfish liver glutamate dehydrogenase showed that the kinetic features of the reaction are very similar with both coenzymes, but reactions with NADH are much faster. The data show that a number of possible mechanisms for the reaction may be discarded, including the compulsory mechanism (previously proposed for the enzyme) in which the sequence of binding is NAD(P)H, NH4+ and 2-oxoglutarate. The kinetic data suggest either a rapid-equilibrium random mechanism or the compulsory mechanism with the binding sequence NH4+, NAD(P)H, 2-oxoglutarate. However, binding studies and protection studies indicate that coenzyme and 2-oxoglutarate do bind to the free enzyme.  相似文献   

9.
Two of the three metabolic subtypes of species utilizing C4-pathway photosynthesis are defined by high activities of either NADP malic enzyme (NADP malic enzyme type) or a coenzyme A (CoA)- and acetyl-CoA-activated NAD malic enzyme (NAD malic enzyme type). These enzymes function to decarboxylate malate as an integral part of the photosynthetic process. Leaves of NADP malic enzyme-type species also contain significant NAD-dependent malic enzyme activity. The purpose of the present study was to examine the nature and photosynthetic role of this activity. With Zea mays, this NAD-dependent activity was found to vary widely in fresh leaf extracts. Incubating extracts at 25 °C resulted in a disproportionate increase in NAD activity so that the final ratio of NADP to NAD activity was always about 5. Strong evidence was provided that the NADP and NAD malic enzyme activities in Z. mays extracts were catalyzed by the same enzyme. These activities remained associated during purification and were coincident after polyacrylamide gel electrophoresis. The pH optimum for NAD-dependent activity was about 7.1, compared with 8.3 for NADP malic enzyme activity. Other properties of the NAD-dependent activity are described, a particularly notable feature being the inhibition of this activity by less than 1 μm NADP and NADPH. Evidence is provided that the NADP malic enzyme of several other NADP malic enzyme-type C4 species also has associated activity toward NAD. We concluded that the NAD-dependent malic enzyme activity would have no significant function in photosynthesis.  相似文献   

10.
1. Changes in the amounts and distribution of protein and respiratory enzymes have been estimated during the life cycle of the fly Lucilia cuprina. 2. The fully fed larva contains about 7mg. of protein, the pupa and newly emerged fly about 4mg., and the mature adult about 3mg. 3. There are two periods of incorporation of protein into particles at the expense of the soluble protein; the first, immediately after pupation, may store protein (0.5mg./insect) for use in adult development; the second, over the period of emergence, was due mainly to the development of the thoracic mitochondria of the adult (0.7mg./insect). 4. In the thorax, cytochrome c oxidase and the dehydrogenases for glycerophosphate, isocitrate (NAD-dependent), succinate and malate appeared initially in small particles (less than 1mu in diameter). 5. In adult development these enzymes were redistributed so that in the mature fly most of the activity was present in larger particles (1-10mu in diameter). 6. During this redistribution the specific activity (mul. of oxygen/hr./mg. of protein) of glycerophosphate dehydrogenase in the small particles was 690 at 1(1/2) days before emergence, 955 at emergence and 980 at 7 days after emergence; the corresponding values for the large particles were 164, 760 and 1220. 7. In the mature fly the highest specific activities (mul. of oxygen/hr./mg. of protein) estimated were: glycerophosphate dehydrogenase 1380, isocitrate dehydrogenase (NAD-dependent and requiring ADP and Mg(2+)) 408, succinate dehydrogenase 122, malate dehydrogenase 190, and cytochrome c oxidase 1360. 8. The results are considered in relation to the development of the flight-muscle sarcosomes.  相似文献   

11.
Cell-free extracts of Methanobacterium thermoautotrophicum were found to contain high activities of the following oxidoreductases (at 60°C): pyruvate dehydrogenase (coenzyme A acetylating), 275 nmol/min per mg of protein; α-ketoglutarate dehydrogenase (coenzyme A acylating), 100 nmol/min per mg; fumarate reductase, 360 nmol/min per mg; malate dehydrogenase, 240 nmol/min per mg; and glyceraldehyde-3-phosphate dehydrogenase, 100 nmol/min per mg. The kinetic properties (apparent Vmax and KM values), pH optimum, temperature dependence of the rate, and specificity for electron acceptors/donors of the different oxidoreductases were examined. Pyruvate dehydrogenase and α-ketoglutarate dehydrogenase were shown to be two separate enzymes specific for factor 420 rather than for nicotinamide adenine dinucleotide (NAD), NADP, or ferredoxin as the electron acceptor. Both activities catalyzed the reduction of methyl viologen with the respective α-ketoacid and a coenzyme A-dependent exchange between the carboxyl group of the α-ketoacid and CO2. The data indicate that the two enzymes are similar to pyruvate synthase and α-ketoglutarate synthase, respectively. Fumarate reductase was found in the soluble cell fraction. This enzyme activity coupled with reduced benzyl viologen as the electron donor, but reduced factor 420, NADH, or NADPH was not effective. The cells did not contain menaquinone, thus excluding this compound as the physiological electron donor for fumarate reduction. NAD was the preferred coenzyme for malate dehydrogenase, whereas NADP was preferred for glyceraldehyde-3-phosphate dehydrogenase. The organism also possessed a factor 420-dependent hydrogenase and a factor 420-linked NADP reductase. The involvement of the described oxidoreductases in cell carbon synthesis is discussed.  相似文献   

12.
In steady-state kinetic studies of ox liver glutamate dehydrogenase in 0.11 M-potassium phosphate buffer, pH7, at 25 degrees C, the concentration of ADP was varied from 0.5 to 1000 microM. Inhibition was observed except when the concentrations of both glutamate and coenzyme were high, when activation was seen. With NAD+ or NADP+ as coenzyme, 200 microM-ADP was sufficient to saturate the enzyme with respect to the major effect of this nucleotide. In the presence of 210 microM-ADP, widely varied concentrations of coenzyme give linear Lineweaver-Burk plots, in marked contrast with results obtained previously for kinetics without ADP. This has allowed evaluation for the reaction with NAD+, NADP+ and acetylpyridine-adenine dinucleotide (315 microM-ADP in the last case) of all four initial rate parameters, i.e. the phi coefficients in the equation: (Formula: see text) where A is coenzyme and B is glutamate. The relative constancy of phi B and of phi AB/phi A with the different coenzymes point to a compulsory-order mechanism with glutamate as the leading substrate. This conclusion, though unexpected, agrees well with various previous observations on the binding of oxidized coenzyme.  相似文献   

13.
T Sanner 《Biochemistry》1975,14(23):5094-5098
The reaction of glutamate dehydrogenase and glutamate (gl) with NAD+ and NADP+ has been studied with stopped-flow techniques. The enzyme was in all experiments present in excess of the coenzyme. The results indicate that the ternary complex (E-NAD(P)H-kg) is present as an intermediate in the formation of the stable complex (E-NAD(P)H-gl). The identification of the complexes is based on their absorption spectra. The binding of the coenzyme to (E-gl) is the rate-limiting step in the formation of (E-NAD(P)H-kg) while the dissociation of alpha-ketoglutarate (kg) from this complex is the rate-limiting step in the formation of (E-NAD(P)H-gl). The Km for glutamate was 20-25 mM in the first reaction and 3 mM in the formation of the stable complex. The Km values were independent of the coenzyme. The reaction rates with NAD+ were approximately 50% greater than those with NADP+. Furthermore, high glutamate concentration inhibited the formation of (E-NADH-kg) while no substrate inhibition was found with NADP+ as coenzyme. ADP enhanced while GTP reduced the rate of (E-NAD(P)H-gl) formation. The rate of formation of (E-NAD(P)H-kg) was inhibited by ADP, while it increased at high glutamate concentration when small amounts of GTP were added. The results show that the higher activity found with NAD+ compared to NADP+ under steady-state assay conditions do not necessarily involve binding of NAD+ to the ADP activating site of the enzyme. Moreover, the substrate inhibition found at high glutamate concentration under steady-state assay condition is not due to the formation of (E-NAD(P)H-gl) as this complex is formed with Km of 3 mM glutamate, and the substrate inhibition is only significant at 20-30 times this concentration.  相似文献   

14.
Previously we found that replacement of seven amino acid residues in a loop region markedly shifted the coenzyme specificity of malate dehydrogenase from NAD(H) toward NADP(H). In the present study, we replaced the seven amino acid residues in the corresponding region of an NAD(H)-dependent lactate dehydrogenase with those of NADP(H)-dependent malate dehydrogenase, and examined the coenzyme specificity of the resulting mutant enzyme. Coenzyme specificity was significantly shifted by 399-fold toward NADPH when k cat/Km(coenzyme) was used as the measure of coenzyme specificity. The effect of the replacements on coenzyme specificity is discussed based on in silico simulation of the three-dimensional structure of the lactate dehydrogenase mutant.  相似文献   

15.
Soluble formate dehydrogenase from Methanobacterium formicicum was purified 71-fold with a yield of 35%. Purification was performed anaerobically in the presence of 10 mM sodium azide which stabilized the enzyme. The purified enzyme reduced, with formate, 50 mumol of methyl viologen per min per mg of protein and 8.2 mumol of coenzyme F420 per min per mg of protein. The apparent Km for 7,8-didemethyl-8-hydroxy-5-deazariboflavin, a hydrolytic derivative of coenzyme F420, was 10-fold greater (63 microM) than for coenzyme F420 (6 microM). The purified enzyme also reduced flavin mononucleotide (Km = 13 microM) and flavin adenine dinucleotide (Km = 25 microM) with formate, but did not reduce NAD+ or NADP+. The reduction of NADP+ with formate required formate dehydrogenase, coenzyme F420, and coenzyme F420:NADP+ oxidoreductase. The formate dehydrogenase had an optimal pH of 7.9 when assayed with the physiological electron acceptor coenzyme F420. The optimal reaction rate occurred at 55 degrees C. The molecular weight was 288,000 as determined by gel filtration. The purified formate dehydrogenase was strongly inhibited by cyanide (Ki = 6 microM), azide (Ki = 39 microM), alpha,alpha-dipyridyl, and 1,10-phenanthroline. Denaturation of the purified formate dehydrogenase with sodium dodecyl sulfate under aerobic conditions revealed a fluorescent compound. Maximal excitation occurred at 385 nm, with minor peaks at 277 and 302 nm. Maximal fluorescence emission occurred at 455 nm.  相似文献   

16.
An advanced strategy of Saccharomyces cerevisiae strain development for fermentation of xylose applies tailored enzymes in the process of metabolic engineering. The coenzyme specificities of the NADPH-preferring xylose reductase (XR) and the NAD?-dependent xylitol dehydrogenase (XDH) have been targeted in previous studies by protein design or evolution with the aim of improving the recycling of NADH or NADPH in their two-step pathway, converting xylose to xylulose. Yeast strains expressing variant pairs of XR and XDH that according to in vitro kinetic data were suggested to be much better matched in coenzyme usage than the corresponding pair of wild-type enzymes, exhibit widely varying capabilities for xylose fermentation. To achieve coherence between enzyme properties and the observed strain performance during fermentation, we explored the published kinetic parameters for wild-type and engineered forms of XR and XDH as possible predictors of xylitol by-product formation (Y(xylitol)) in yeast physiology. We found that the ratio of enzymatic reaction rates using NADP(H) and NAD(H) that was calculated by applying intracellular reactant concentrations to rate equations derived from bi-substrate kinetic analysis, succeeded in giving a statistically reliable forecast of the trend effect on Y(xylitol). Prediction based solely on catalytic efficiencies with or without binding affinities for NADP(H) and NAD(H) were not dependable, and we define a minimum demand on the enzyme kinetic characterization to be performed for this purpose. An immediate explanation is provided for the typically lower Y(xylitol) in the current strains harboring XR engineered for utilization of NADH as compared to strains harboring XDH engineered for utilization of NADP?. The known XDH enzymes all exhibit a relatively high K(m) for NADP? so that physiological boundary conditions are somewhat unfavorable for xylitol oxidation by NADP?. A criterion of physiological fitness is developed for engineered XR working together with wild-type XDH.  相似文献   

17.
Kallberg Y  Persson B 《The FEBS journal》2006,273(6):1177-1184
Dehydrogenases and reductases are enzymes of fundamental metabolic importance that often adopt a specific structure known as the Rossmann fold. This fold, consisting of a six-stranded beta-sheet surrounded by alpha-helices, is responsible for coenzyme binding. We have developed a method to identify Rossmann folds and predict their coenzyme specificity (NAD, NADP or FAD) using only the amino acid sequence as input. The method is based upon hidden Markov models and sequence pattern analysis. The prediction sensitivity is 79% and the selectivity close to 100%. The method was applied on a set of 68 genomes, representing the three kingdoms archaea, bacteria and eukaryota. In prokaryotes, 3% of the genes were found to code for Rossmann-fold proteins, while the corresponding ratio in eukaryotes is only around 1%. In all genomes, NAD is the most preferred cofactor (41-49%), followed by NADP with 30-38%, while FAD is the least preferred cofactor (21%). However, the NAD preponderance over NADP is most pronounced in archaea, and least in eukaryotes. In all three kingdoms, only 3-8% of the Rossmann proteins are predicted to have more than one membrane-spanning segment, which is much lower than the frequency of membrane proteins in general. Analysis of the major protein types in eukaryotes reveals that the most common type (26%) of the Rossmann proteins are short-chain dehydrogenases/reductases. In addition, the identified Rossmann proteins were analyzed with respect to further protein types, enzyme classes and redundancy. The described method is available at http://www.ifm.liu.se/bioinfo, where the preferred coenzyme and its binding region are predicted given an amino acid sequence as input.  相似文献   

18.
Gastric tissues from amphibian Rana perezi express the only vertebrate alcohol dehydrogenase (ADH8) that is specific for NADP(H) instead of NAD(H). In the crystallographic ADH8-NADP+ complex, a binding pocket for the extra phosphate group of coenzyme is formed by ADH8-specific residues Gly223-Thr224-His225, and the highly conserved Leu200 and Lys228. To investigate the minimal structural determinants for coenzyme specificity, several ADH8 mutants involving residues 223 to 225 were engineered and kinetically characterized. Computer-assisted modeling of the docked coenzymes was also performed with the mutant enzymes and compared with the wild-type crystallographic binary complex. The G223D mutant, having a negative charge in the phosphate-binding site, still preferred NADP(H) over NAD(H), as did the T224I and H225N mutants. Catalytic efficiency with NADP(H) dropped dramatically in the double mutants, G223D/T224I and T224I/H225N, and in the triple mutant, G223D/T224I/H225N (kcat/KmNADPH = 760 mm-1 min-1), as compared with the wild-type enzyme (kcat/KmNADPH = 133330 mm-1 min-1). This was associated with a lower binding affinity for NADP+ and a change in the rate-limiting step. Conversely, in the triple mutant, catalytic efficiency with NAD(H) increased, reaching values (kcat/KmNADH = 155000 mm-1 min-1) similar to those of the wild-type enzyme with NADP(H). The complete reversal of ADH8 coenzyme specificity was therefore attained by the substitution of only three consecutive residues in the phosphate-binding site, an unprecedented achievement within the ADH family.  相似文献   

19.
1. Skeletal muscle mitochondrial NAD(P)-dependent malic enzyme [EC 1.1.1. 39, L-malate:NAD+ oxidoreductase (decarboxylating)] from herring could use both coenzymes, NAD and NADP, in a similar manner. 2. The coenzyme preference of mitochondrial NAD(P)-dependent malic enzyme was probed using dual wavelength spectroscopy and pairing the natural coenzymes, NAD or NADP with their respective thionicotinamide analogues, s-NADP or s-NAD, that have absorbance maxima in reduced forms at 400 nm. 3. s-NAD and s-NADP were found to be good alternate substrates for NAD(P)-dependent malic enzyme, the apparent Km values for the thioderivatives were similar to those of the corresponding natural coenzymes. 4. ATP produced greater inhibition of the NAD or s-NAD linked reactions than of the NADP or s-NADP-linked reactions of skeletal muscle mitochondrial NAD(P)-dependent malic enzyme. 5. At 5 mM malate concentration and in the presence of 2 mM ATP the NADP-linked reaction is favoured and the activity ratios, V(s-NADP)/V(NAD) or V(NADP)/V(s-NAD), are 6 and 26, respectively.  相似文献   

20.
SYNOPSIS. Nicotinamide adenine dinucleotide phosphate (NADP) and nicotinamide adenine dinucleotide (NAD) linked isocitrate dehydrogenase and NADP linked alcohol dehydrogenase have been detected in Euglena gracilis var. bacillaris. The NADP isocitrate dehydrogenase showed half-maximal activity at a concentration of 3 × 10?5 M DL-isocitrate, but did not follow simple Michaelis-Menten kinetics with respect to substrate concentration. The optimal NADP concentration was about 0.06 mM, and activity fell off sharply on either side of this optimum. Fresh preparations of the enzyme migrated as single bands in disc electrophoresis, but two enzymatically active bands were present after frozen storage. The NAD isocitrate dehydrogenase followed Michaelis-Menten kinetics with respect to substrate. In crude extracts, no requirement for adenosine monophosphate, adenosine diphosphate, or sulfhydryl compounds could be found. NADP alcohol dehydrogenase activity could be found with either ethanol or propanol as substrate. Low concentrations of coenzyme A were moderately inhibitory. In tris(hydroxymethyl) aminomethane buffer (tris buffer), Euglena extracts reduced NAD slowly in the absence of exogenous substrate. In the absence of tris, no such reduction occurred. A similar phenomenon was observed with NADP.  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