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 62 毫秒
1.
Abstract The facultatively methylotrophic bacterium Pseudomonas sp. 101, grown on methanol in presence of molybdate, contains a new formate dehydrogenase (N-FDH) catalyzing NAD+-dependent oxidation of formate. The activity of this N-FDH could also be measured in presence of artificial electron acceptors, ferricyanide and 2,6-dichlorophenol indophenol. This new enzyme is absent in cells grown on a methanol-containing medium with tungstate, where only another two, previously described formate dehydrogenases, which are active only with NAD+ or only with artificial acceptors, respectively, were determined. The N-FDH was partially purified by a combination of ion-exchange and gel-filtration chromatography, and was shown to differ in its properties from the known NAD+-dependent counterpart.  相似文献   

2.
Organoautotrophic growth of Alcaligenes eutrophus on formate was dependent on the presence of molybdate in the medium. Supplementation of the medium with tungstate lead to growth cessation. Corresponding effects of these anions were observed for the activity of the soluble, NAD(+)-linked formate dehydrogenase (S-FDH; EC 1.2.1.2) of the organism. Lack of molybdate or presence of tungstate resulted in an almost complete loss of S-FDH activity. S-FDH was purified to near homogeneity in the presence of nitrate as a stabilizing agent. The native enzyme exhibited an M(r) of 197,000 and a heterotetrameric quaternary structure with nonidentical subunits of M(r) 110,000 (alpha), 57,000 (beta), 19,400 (gamma), and 11,600 (delta). It contained 0.64 g-atom of molybdenum, 25 g-atom of nonheme iron, 20 g-atom of acid-labile sulfur, and 0.9 mol of flavin mononucleotide per mol. The fluorescence spectrum of iodine-oxidized S-FDH was nearly identical to the form A spectrum of milk xanthine oxidase, proving the presence of a pterin cofactor. The molybdenum-complexing cofactor was identified as molybdopterin guanine dinucleotide in an amount of 0.71 mol/mol of S-FDH. Apparent Km values of 3.3 mM for formate and 0.09 mM for NAD+ were determined. The enzyme coupled the oxidation of formate to a number of artificial electron acceptors and was strongly inactivated by formate in the absence of NAD+. It was inhibited by cyanide, azide, nitrate, and Hg2+ ions. Thus, the enzyme belongs to a new group of complex molybdo-flavo Fe-S FDH that so far has been detected in only one other aerobic bacterium.  相似文献   

3.
Mycobacterium vaccae 10 growing in methanol medium synthesizes two inducible alternative NAD(+)-dependent formate dehydrogenases (FDH). In the presence of molybdenum, the dominating form of the enzyme is FDHI with Mr 440 kDa and Km 0.32 mM for sodium formate. FDHI reduced ferricyanide as well as NAD+, and it was reversibly inactivated by formate. NAD+ stabilized FDHI against this inactivation. Under conditions of artificial molybdenum deficiency (tungsten in the medium), the second enzyme (FDHII) appeared with Mr about 93 kDa and Km 8.3 mM for sodium formate, and no FDHI activity was detected. FDHII did not reduce ferricyanide and was not inactivated by formate. The activity of FDHI was restored in tungsten-grown cells by pulse addition of molybdenum under conditions of blocked protein synthesis, suggesting the pre-existence of inactive apo-FDHI.  相似文献   

4.
Extracts of aerobically, CO-autotrophically grown cells of Pseudomonas carboxydovorans were shown to catalyze the oxidation of CO to CO(2) in the presence of methylene blue, pyocyanine, thionine, phenazine methosulfate, or toluylene blue under strictly anaerobic conditions. Viologen dyes and NAD(P)(+) were ineffective as electron acceptors. The same extracts catalyzed the oxidation of formate and of hydrogen gas; the spectrum of electron acceptors was identical for the three substrates, CO, formate, and H(2). The CO- and the formate-oxidizing activities were found to be soluble enzymes, whereas hydrogenase was membrane bound exclusively. The rates of oxidation of CO, formate, and H(2) were measured spectrophotometrically following the reduction of methylene blue. The rate of carbon monoxide oxidation followed simple Michaelis-Menten kinetics; the apparent K(m) for CO was 45 muM. The reaction rate was maximal at pH 7.0, and the temperature dependence followed the Arrhenius equation with an activation energy (DeltaH(0)) of 35.9 kJ/mol (8.6 kcal/mol). Neither free formate nor hydrogen gas is an intermediate of the CO oxidation reaction. This conclusion is based on the differential sensitivity of the activities of formate dehydrogenase, hydrogenase, and CO dehydrogenase to heat, hypophosphite, chlorate, cyanide, azide, and fluoride as well as on the failure to trap free formate or hydrogen gas in coupled optical assays. These results support the following equation for CO oxidation in P. carboxydovorans: CO + H(2)O --> CO(2) + 2 H(+) + 2e(-) The CO-oxidizing activity of P. carboxydovorans differed from that of Clostridium pasteurianum by not reducing viologen dyes and by a pH optimum curve that did not show an inflection point.  相似文献   

5.
Glyceraldehyde-3-phosphate dehydrogenase (GAPDH), a key enzyme of central carbon metabolism, was studied in a Bacillus cereus strain isolated from the phosphate layer from Morocco. Enzymatic assays with cell extracts demonstrated that when grown on Luria-Bertani (LB) medium, B. cereus contains a major NAD+-dependent GAPDH activity and only traces of NADP+-dependent activity, but in cells grown on Pi-supplemented LB medium a strong increase of the NADP+-dependent activity, that became predominant, occurs concurrently with a GAPDH protein increase. Our results show that B. cereus possesses two GAPDH activities, namely NAD+- and NADP+-dependent, catalyzed by two enzymes with distinct coenzyme specificity and different phosphate regulation patterns. The finding of a phosphate-stimulated NADP+-dependent GAPDH in B. cereus indicates that this bacterium can modulate its primary carbon metabolism according to phosphate availability.  相似文献   

6.
Formate dehydrogenase (EC 1.2.1.2) prepared from peas (Pisum sativum) was a two-subunit enzyme. The enzyme accelerated the formation of an NAD+-cyanide compound having an adsorption band at 330 nm. The enzyme was able to bind one NAD+ molecule per each subunit but only 1 mole of NAD+-cyanide compound was formed per two subunits. The complex of NAD+, cyanide, and the enzyme was very stable and had no catalytic activity. Azide inhibited the formate dehydrogenase reaction in two different ways. By incubation of the enzyme with azide in the presence of NAD+, half of its catalytic activity was lost. The remaining activity was also inhibited by azide but this inhibition was removed competively by formate. Contrary to the case of cyanide the inhibition by azide could be removed by dialysis and no spectral species due to the addition compound of NAD+ and azide could be observed. The data from double recipricol plots of the initial velocity and the formate concentration led to a conclusion that formate dehydrogenase has two sites with about equal catalytic activity. The Km for formate was different for the two catalytic sites (1.67 and 6.25 mM) but the difference was not noticeable in the case of the Km for NAD+.  相似文献   

7.
It is generally known that cofactors play a major role in the production of different fermentation products. This paper is part of a systematic study that investigates the potential of cofactor manipulations as a new tool for metabolic engineering. The NADH/NAD+ cofactor pair plays a major role in microbial catabolism, in which a carbon source, such as glucose, is oxidized using NAD+ and producing reducing equivalents in the form of NADH. It is crucially important for continued cell growth that NADH be oxidized to NAD+ and a redox balance be achieved. Under aerobic growth, oxygen is used as the final electron acceptor. While under anaerobic growth, and in the absence of an alternate oxidizing agent, the regeneration of NAD+ is achieved through fermentation by using NADH to reduce metabolic intermediates. Therefore, an increase in the availability of NADH is expected to have an effect on the metabolic distribution. We have previously investigated a genetic means of increasing the availability of intracellular NADH in vivo by regenerating NADH through the heterologous expression of an NAD(+)-dependent formate dehydrogenase and have demonstrated that this manipulation provoked a significant change in the final metabolite concentration pattern both anaerobically and aerobically (Berríos-Rivera et al., 2002, Metabolic engineering of Escherichia coli: increase of NADH availability by overexpressing an NAD(+)-dependent formate dehydrogenase, Metabolic Eng. 4, 217-229). The current work explores further the effect of substituting the native cofactor-independent formate dehydrogenase (FDH) by an NAD(+)-dependent FDH from Candida boidinii on the NAD(H/+) levels, NADH/NAD+ ratio, metabolic fluxes and carbon-mole yields in Escherichia coli under anaerobic chemostat conditions. Overexpression of the NAD(+)-dependent FDH provoked a significant redistribution of both metabolic fluxes and carbon-mole yields. Under anaerobic chemostat conditions, NADH availability increased from 2 to 3 mol NADH/mol glucose consumed and the production of more reduced metabolites was favored, as evidenced by a dramatic increase in the ethanol to acetate ratio and a decrease in the flux to lactate. It was also found that the NADH/NAD+ ratio should not be used as a sole indicator of the oxidation state of the cell. Instead, the metabolic distribution, like the Et/Ac ratio, should also be considered because the turnover of NADH can be fast in an effort to achieve a redox balance.  相似文献   

8.
NAD+-dependent formate dehydrogenase was screened in various bacterial strains. Facultative methanol-utilizing bacteria isolated from soil samples, acclimated to a medium containing methanol and formate at pH 9.5, were classified as members of the genus Moraxella. From a crude extract of Moraxella sp. strain C-1, formate dehydrogenase was purified to homogeneity, as judged by disc gel electrophoresis. The enzyme has an isoelectric point of 3.9 and a molecular weight of approximately 98,000. The enzyme is composed of two identical subunits with molecular weights of about 48,000. The apparent Km values for sodium formate and NAD+ were calculated to be 13 mM and 0.068 mM, respectively.  相似文献   

9.
Cell extracts of acetate-grown Methanosarcina strain TM-1 and Methanosarcina acetivorans both contained CH3-S-CoM methylreductase activity. The methylreductase activity was supported by CO and H2 but not by formate as electron donors. The CO-dependent activity was equivalent to the H2-dependent activity in strain TM-1 and was fivefold higher than the H2-dependent activity of M. acetivorans. When strain TM-1 was cultured on methanol, the CO-dependent activity was reduced to 5% of the activity in acetate-grown cells. Methanobacterium formicicum grown on H2-CO2 contained no CO-dependent methylreductase activity. The CO-dependent methylreductase of strain TM-1 had a pH optimum of 5.5 and a temperature optimum of 60 degrees C. The activity was stimulated by the addition of MgCl2 and ATP. Both acetate-grown strain TM-1 and acetate-grown M. acetivorans contained CO dehydrogenase activities of 9.1 and 3.8 U/mg, respectively, when assayed with methyl viologen. The CO dehydrogenase of acetate-grown cells rapidly reduced FMN and FAD, but coenzyme F420 and NADP+ were poor electron acceptors. No formate dehydrogenase was detected in either organism when grown on acetate. The results suggest that a CO-dependent CH3-S-CoM methylreductase system is involved in the pathway of the conversion of acetate to methane and that free formate is not an intermediate in the pathway.  相似文献   

10.
Summary Formate dehydrogenase in extracts of the facultative phototroph, Rhodopseudomonas palustris was shown to be soluble and NAD-linked. The flavin nucleotides, FMN and FAD, stimulated the rate of NAD reduction about fourfold. Reduction of artificial electron acceptors such as DCPIP and cytochrome c was also stimulated by FMN and FAD. The pH optimum for the reduction of NAD was pH 8.0, in contrast to pH 6.8 for cytochrome c and DCPIP reduction. The apparent K m for formate as measured by NAD reduction was 2.6×10-4 M. Although the addition of thiosulfate or yeast extract to the formate medium increased both the growth rate and yield of Rhps. palustris, they had little effect on the activity of formate dehydrogenase.  相似文献   

11.
Properties of recombinant wild type (WT) and six-histidine tag-fused (His(6)) putidaredoxin reductase (Pdr), a FAD-containing component of the soluble cytochrome P450cam monooxygenase system from Pseudomonas putida, have been studied. Both WT and His(6) Pdr were found to undergo a monomer-dimer association-dissociation and were partially present as an NAD(+)-bound form. Although molecular, spectral, and electron transferring properties of recombinant His(6) Pdr to artificial and native electron acceptors were similar to those of the WT protein, the presence of eight additional C-terminal amino acid residues, Pro-Arg-His-His-His-His-His-His, had a crucial effect on the enzyme interaction with oxidized pyridine nucleotide. Under anaerobic conditions, NAD(+) induced in His(6) Pdr spectral changes indicative of flavin reduction and formation of the charge transfer complex between the reduced FAD and NAD(+). The reaction proceeded considerably faster in the presence of free histidine and thiol-reducing agents, such as dithiothreitol and reduced glutathione. In the presence of any of these three reagents, NAD(+) was capable of inducing reduction of the flavin in WT Pdr. Free thiol groups were identified as an internal source of electrons in the enzyme. The results showed that WT and His(6) Pdr were able to function as NAD(H)-dependent dithiol/disulfide oxidoreductases catalyzing both forward and reverse reactions, NAD(+)-dependent oxidation of thiols, and NADH-dependent reduction of disulfides. This function of the flavoprotein can be dissociated from electron transfer to putidaredoxin. Similarity of Pdr to the enzymes of the glutathione reductase family is discussed.  相似文献   

12.
Escherichia coli overexpressing a NAD(+)-dependent formate dehydrogenase (FDH) from Candida boidinii was grown in chemostat culture on various carbon sources at 0.05 h(-1) dilution rate, under anaerobic conditions using defined medium and compared to a control without the heterologous FDH pathway. Metabolic fluxes, NADH/NAD(+) ratios and NAD(H/(+)) levels were determined under a range of intracellular NADH availability. The effect of NADH manipulation on the distribution of metabolic fluxes in E. coli was assessed under steady-state conditions. The heterologous FDH pathway converts 1 mol of formate into 1 mol of NADH and carbon dioxide, in contrast with the native FDH where no cofactor involvement is present. Previously, we found that this NADH regeneration system doubled the maximum yield of NADH from 2 to 4 mol NADH/mol glucose consumed and reached 4.6 mol NADH/mol of substrate when sorbitol was used as a carbon source in a complex medium. In the current study, it was found that higher NADH yields and NADH/NAD(+) ratios were achieved with our in vivo NADH regeneration system compared to a control lacking the new FDH pathway in the three carbon sources (glucose, gluconate and sorbitol) examined suggesting a more reduced intracellular environment. The total NAD(H/(+)) amounts were very similar for all the combinations studied. It was also found that the ethanol to acetate ratio increased with increased NADH availability. This ratio increased from 1.05 for the control strain in glucose to 9.45 for the strain expressing the heterologous NAD(+)-dependent FDH in sorbitol.  相似文献   

13.
Microbiology - Comparative phylogenetic analysis of NAD+-dependent formate dehydrogenases (NAD+-FDH) genes, which have been detected in all available genomes of methylotrophs of the genera...  相似文献   

14.
The expression of the recombinant wild-type NAD+- and mutant NADP+-dependent formate dehydrogenases (EC 1.2.1.2., FDH) from the methanol-utilizing bacterium Pseudomonas sp. 101 in Escherichia coli cells has been improved to produce active and soluble enzyme up to the level of 50% of total soluble proteins. The cultivation process for E. coli/pFDH8a and E. coli/pFDH8aNP cells was optimized and scaled up to a volume of 100 L. A downstream purification process has been developed to produce technical grade NAD+- and NADP+-specific formate dehydrogenases in pilot scale, utilizing extraction in aqueous two-phase systems.  相似文献   

15.
Both the external oxidation of NADH and NADPH in intact potato (Solanum tuberosum L. cv. Bintje) tuber mitochondria and the rotenone-insensitive internal oxidation of NADPH by inside-out submitochondrial particles were dependent on Ca2+. The stimulation was not due to increased permeability of the inner mitochondrial membrane. Neither the membrane potential nor the latencies of NAD(+)-dependent and NADP(+)-dependent malate dehydrogenases were affected by the addition of Ca2+. The pH dependence and kinetics of Ca(2+)-dependent NADPH oxidation by inside-out submitochondrial particles were studied using three different electron acceptors: O2, duroquinone and ferricyanide. Ca2+ increased the activity with all acceptors with a maximum at neutral pH and an additional minor peak at pH 5.8 with O2 and duroquinone. Without Ca2+, the activity was maximal around pH 6. The Km for NADPH was decreased fourfold with ferricyanide and duroquinone, and twofold with O2 as acceptor, upon addition of Ca2+. The Vmax was not changed with ferricyanide as acceptor, but increased twofold with both duroquinone and O2. Half-maximal stimulation of the NADPH oxidation was found at 3 microM free Ca2+ with both O2 and duroquinone as acceptors. This is the first report of a membrane-bound enzyme inside the inner mitochondrial membrane which is directly dependent on micromolar concentrations of Ca2+. Mersalyl and dicumarol, two potent inhibitors of the external NADH dehydrogenase in plant mitochondria, were found to inhibit internal rotenone-insensitive NAD(P)H oxidation, at the same concentrations and in manners very similar to their effects on the external NAD(P)H oxidation.  相似文献   

16.
Whole cells of Pseudomonas putida N.C.I.B 9869, when grown on either 3,5-xylenol or p-cresol, oxidized both m- and p-hydroxybenzyl alcohols. Two distinct NAD+-dependent m-hydroxybenzyl alcohol dehydrogenases were purified from cells grown on 3,5-xylenol. Each is active with a range of aromatic alcohols, including both m- and p-hydroxybenzyl alcohol, but differ in their relative rates with the various substrates. An NAD+-dependent alcohol dehydrogenase was also partially purified from p-cresol grown cells. This too was active with m- and p-hydroxybenzyl alcohol and other aromatic alcohols, but was not identical with either of the other two dehydrogenases. All three enzymes were unstable, but were stabilized by dithiothreitol and all were inhibited with p-chloromercuribenzoate. All were specific for NAD+ and each was shown to catalyse conversion of alcohol into aldehyde.  相似文献   

17.
The primary kinetic isotope effect of the reaction catalyzed by NAD+-dependent formate dehydrogenase (EC 1.2.1.2.) from the methylotrophic bacterium Pseudomonas sp. 101 has been studied. Analysis of the ratios HVm/DVm and H(Vm/KM)/D(Vm/KM) in the pH range 6.1-7.9 showed that the transfer of hydride ion in ternary enzyme-substrate complex is a limiting step of the reaction, and the formate binding to the binary complex (formate dehydrogenase + NAD+) reached equilibrium when the pH of the medium was increased. An approach has been developed to determine the elementary constants of substrate association (kon) and dissociation (koff) at the stages of the binary--ternary enzyme-substrate complexes for the random equilibrium 2-substrate kinetic mechanism. The kon and koff values obtained for the bacterial formate dehydrogenase by using the proposed approach for NAD+ were (4.8 +/- 0.8)*10(5)M-1s-1 and (90 +/- 10) s-1, and for formate (2.0 +/- 1.0)*10(4) M-1s-1 and (60 +/- 20) s-1, respectively.  相似文献   

18.
Metabolic engineering studies have generally focused on manipulating enzyme levels through either the amplification, addition, or deletion of a particular pathway. However, with cofactor-dependent production systems, once the enzyme levels are no longer limiting, cofactor availability and the ratio of the reduced to oxidized form of the cofactor can become limiting. Under these situations, cofactor manipulation may become crucial in order to further increase system productivity. Although it is generally known that cofactors play a major role in the production of different fermentation products, their role has not been thoroughly and systematically studied. However, cofactor manipulations can potentially become a powerful tool for metabolic engineering. Nicotinamide adenine dinucleotide (NAD) functions as a cofactor in over 300 oxidation-reduction reactions and regulates various enzymes and genetic processes. The NADH/NAD+ cofactor pair plays a major role in microbial catabolism, in which a carbon source, such as glucose, is oxidized using NAD+ producing reducing equivalents in the form of NADH. It is crucially important for continued cell growth that NADH be oxidized to NAD+ and a redox balance be achieved. Under aerobic growth, oxygen is used as the final electron acceptor. While under anaerobic growth, and in the absence of an alternate oxidizing agent, the regeneration of NAD+ is achieved through fermentation by using NADH to reduce metabolic intermediates. Therefore, an increase in the availability of NADH is expected to have an effect on the metabolic distribution. This paper investigates a genetic means of manipulating the availability of intracellular NADH in vivo by regenerating NADH through the heterologous expression of an NAD(+)-dependent formate dehydrogenase. More specifically, it explores the effect on the metabolic patterns in Escherichia coli under anaerobic and aerobic conditions of substituting the native cofactor-independent formate dehydrogenase (FDH) by and NAD(+)-dependent FDH from Candida boidinii. The over-expression of the NAD(+)-dependent FDH doubled the maximum yield of NADH from 2 to 4 mol NADH/mol glucose consumed, increased the final cell density, and provoked a significant change in the final metabolite concentration pattern both anaerobically and aerobically. Under anaerobic conditions, the production of more reduced metabolites was favored, as evidenced by a dramatic increase in the ethanol-to-acetate ratio. Even more interesting is the observation that during aerobic growth, the increased availability of NADH induced a shift to fermentation even in the presence of oxygen by stimulating pathways that are normally inactive under these conditions.  相似文献   

19.
Extracts of Pseudomonas C grown on methanol as a sole carbon and energy source contain a methanol dehydrogenase activity which can be coupled to phenazine methosulfate. This enzyme catalyzes two reactions namely the conversion of methanol to formaldehyde (phenazine methosulfate coupled) and the oxidation of formaldehyde to formate (2,6-dichloroindophenol-coupled). Activities of glutathione-dependent formaldehyde dehydrogenase (NAD+) and formate dehydrogenase (NAD+) were also detected in the extracts. The addition of D-ribulose 5-phosphate to the reaction mixtures caused a marked increase in the formaldehyde-dependent reduction of NAD+ or NADP+. In addition, the oxidation of [14C]formaldehyde to CO2, by extracts of Pseudomonas C, increased when D-ribulose 5-phosphate was present in the assay mixtures. The amount of radioactivity found in CO2, was 6;8-times higher when extracts of methanol-grown Pseudomonas C were incubated for a short period of time with [1-14C]glucose 6-phosphate than with [U-14C]glucose 6-phosphate. These data, and the presence of high specific activities of hexulose phosphate synthase, phosphoglucoisomerase, glucose-6-phosphate dehydrogenase and 6-phosphogluconate dehydrogenase indicate that in methanol-grown Pseudomonas C, formaldehyde carbon is oxidized to CO2 both via a cyclic pathway which includes the enzymes mentioned and via formate as an oxidation intermediate, with the former predominant.  相似文献   

20.
NAD+-dependent formate dehydrogenase (FDH, EC 1.2.1.2) is one of the best enzymes for the purpose of NADH regeneration in dehydrogenase-based synthesis of optically active compounds. Low operational stability and high production cost of native FDHs limit their application in commercial production of chiral compounds. The review summarizes the results on engineering of bacterial and yeast FDHs aimed at improving their chemical and thermal stability, catalytic activity, switch in coenzyme specificity from NAD+ to NADP+ and overexpression in Escherichia coli cells.  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