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.
1. A cell-free system from Pseudomonas fluorescens catalysed the oxidative demethylation and subsequent ring-cleavage of vanillate, with uptake of 2·5 moles of oxygen/mole of substrate. 2. Demethylation involved absorption of 0·5 mole of oxygen/mole, and required reduced glutathione (GSH) and nucleotide (probably NADPH) as cofactors, with further possible requirements, the natures of which are discussed. 3. Incomplete evidence suggested that the aromatic ring was opened via protocatechuate and the appropriate oxygenase, with absorption of 1 mole of oxygen/mole of substrate, eventually yielding β-oxoadipate. 4. The methyl group was removed sequentially as formaldehyde, formate and carbon dioxide, the steps catalysed respectively by formaldehyde dehydrogenase, which required GSH and NAD+, and formate dehydrogenase. Each enzyme was cytochrome-linked and accounted for absorption of 0·5mole of oxygen/mole of substrate. 5. All enzymes except formate dehydrogenase, which was a cell-wall enzyme, resided in the soluble fraction of the extract. The demethylase could not be resolved because of unknown cofactor requirements.  相似文献   

2.
The oxidation of one carbon compounds (methane, methanol, formaldehyde, formate) and primary alcohols (ethanol, propanol, butanol) supported the assimilation of [1-14C]acetate by cell suspensions of type I obligate methylotroph; Pseudomonas methanica, Texas strain, and type II obligate methylotroph, Methylosinus trichosporium, strain PG. The amount of oxygen consumed and substrate oxidized correlated with the amount of [1-14C]acetate assimilated during oxidation of C-1 compounds and primary alcohols.Oxidation of methanol, formaldehyde, and primary alcohols in extracts of Pseudomonas methanica, Texas strain, and Methylosinus trichosporium, strain PG, was catalyzed by a phenazine methosulfate linked, ammonium ion dependent methanol dehydrogenase. The oxidation of aldehydes was catalyzed by a phenazine methosulfate linked, ammonium ion independent aldehyde dehydrogenase. Formate was oxidized by a NAD+ linked formate dehydrogenase.Deceased.This work was supported by Grant GB 8173 from the National Science Foundation and by a grant from the Robert A. Welch Foundation.  相似文献   

3.
A whole-cell biotransformation system for the reduction of prochiral carbonyl compounds, such as methyl acetoacetate, to chiral hydroxy acid derivatives [methyl (R)-3-hydroxy butanoate] was developed in Escherichia coli by construction of a recombinant oxidation/reduction cycle. Alcohol dehydrogenase from Lactobacillus brevis catalyzes a highly regioselective and enantioselective reduction of several ketones or keto acid derivatives to chiral alcohols or hydroxy acid esters. The adh gene encoding for the alcohol dehydrogenase of L. brevis was expressed in E. coli. As expected, whole cells of the recombinant strain produced only low quantities of methyl (R)-3-hydroxy butanoate from the substrate methyl acetoacetate. Therefore, the fdh gene from Mycobacterium vaccae N10, encoding NAD+-dependent formate dehydrogenase, was functionally coexpressed. The resulting two-fold recombinant strain exhibited an in vitro catalytic alcohol dehydrogenase activity of 6.5 units mg–1 protein in reducing methyl acetoacetate to methyl (R)-3-hydroxy butanoate with NADPH as the cofactor and 0.7 units mg–1 protein with NADH. The in vitro formate dehydrogenase activity was 1.3 units mg–1 protein. Whole resting cells of this strain catalyzed the formation of 40 mM methyl (R)-3-hydroxy butanoate from methyl acetoacetate. The product yield was 100 mol% at a productivity of 200 mol g–1 (cell dry weight) min–1. In the presence of formate, the intracellular [NADH]/[NAD+] ratio of the cells increased seven-fold. Thus, the functional overexpression of alcohol dehydrogenase in the presence of formate dehydrogenase was sufficient to enable and sustain the desired reduction reaction via the relatively low specific activity of alcohol dehydrogenase with NADH, instead of NADPH, as a cofactor.  相似文献   

4.
A cytochrome P450BM3‐catalyzed reaction system linked by a two‐step cofactor regeneration was investigated in a cell‐free system. The two‐step cofactor regeneration of redox cofactors, NADH and NADPH, was constructed by NAD+‐dependent bacterial glycerol dehydrogenase (GLD) and bacterial soluble transhydrogenase (STH) both from Escherichia coli. In the present system, the reduced cofactor (NADH) was regenerated by GLD from the oxidized cofactor (NAD+) using glycerol as a sacrificial cosubstrate. The reducing equivalents were subsequently transferred to NADP+ by STH as a cycling catalyst. The resultant regenerated NADPH was used for the substrate oxidation catalyzed by cytochrome P450BM3. The initial rate of the P450BM3‐catalyzed reaction linked by the two‐step cofactor regeneration showed a slight increase (approximately twice) when increasing the GLD units 10‐fold under initial reaction conditions. In contrast, a 10‐fold increase in STH units resulted in about a 9‐fold increase in the initial reaction rate, implying that transhydrogenation catalyzed by STH was the rate‐determining step. In the system lacking the two‐step cofactor regeneration, 34% conversion of 50 μM of a model substrate (p‐nitrophenoxydecanoic acid) was attained using 50 μM NADPH. In contrast, with the two‐step cofactor regeneration, the same amount of substrate was completely converted using 5 μM of oxidized cofactors (NAD+ and NADP+) within 1 h. Furthermore, a 10‐fold dilution of the oxidized cofactors still led to approximately 20% conversion in 1 h. These results indicate the potential of the combination of GLD and STH for use in redox cofactor recycling with catalytic quantities of NAD+ and NADP+. © 2009 American Institute of Chemical Engineers Biotechnol. Prog., 2009  相似文献   

5.
The in vivo flux through the oxidative branch of the pentose phosphate pathway (oxPPP) in Penicillium chrysogenum was determined during growth in glucose/ethanol carbon-limited chemostat cultures, at the same growth rate. Non-stationary 13C flux analysis was used to measure the oxPPP flux. A nearly constant oxPPP flux was found for all glucose/ethanol ratios studied. This indicates that the cytosolic NADPH supply is independent of the amount of assimilated ethanol. The cofactor assignment in the model of van Gulik et al. (Biotechnol Bioeng 68(6):602–618, 2000) was supported using the published genome annotation of P. chrysogenum. Metabolic flux analysis showed that NADPH requirements in the cytosol remain nearly the same in these experiments due to constant biomass growth. Based on the cytosolic NADPH balance, it is known that the cytosolic aldehyde dehydrogenase in P. chrysogenum is NAD +  dependent. Metabolic modeling shows that changing the NAD + -aldehyde dehydrogenase to NADP + -aldehyde dehydrogenase can increase the penicillin yield on substrate.  相似文献   

6.
An NAD+-linked, reduced glutathione-dependent formaldehyde dehydrogenase was purified to homogeneity from soluble extracts of methanol-grown yeast, Pichia sp. Formaldehyde and methylglyoxal are oxidized in the presence of NAD+ as an electron acceptor. NADP+ could not replace NAD+. Other straight chain aldehydes (C2–C6 tested), branched-chain aldehydes (e.g., isobutyaldehyde), aromatic aldehydes (e.g., salicylal-dehyde, benzaldehyde), glutyraldehyde, glyceraldehyde, glycoaldehyde, and glyoxal-dehyde tested were not oxidized by the purified formaldehyde dehydrogenase. The product of formaldehyde oxidation by purified enzyme was demonstrated to be S-for-mylglutathione by measuring the absorption at 240 nm due to the formation of thioester of formaldehyde and reduced glutathione. The Km values for NAD+, formaldehyde, and reduced glutathione were 0.12, 0.31, and 0.16 mm, respectively, for the forward reaction at pH 8.0. The purified formaldehyde dehydrogenase also catalyzed the reduction of S-formylglutathione in the presence of NADH. Formate was not reduced by the purified enzyme. The Km values for S-formylglutathione and NADH were 0.60 and 0.25 mm, respectively, for the reverse reaction at pH 6.0. Formaldehyde dehydrogenase has a molecular weight of 84,000 as determined by gel filtration and subunit molecular weight of 41,000 as determined by sodium dodecyl sulfate-gel electrophoresis. S-Formylglutathione, a product of formaldehyde oxidation, was oxidized by the partially purified formate dehydrogenase from Pichia sp. Formate dehydrogenase has a higher affinity toward S-formylglutathione (Km value 1.8 mm) than toward formate (Km value 25 mm). Antiserum prepared against the purified formaldehyde dehydrogenase from Pichia sp. NRRL-Y-11328 forms strong precipitin bands with isofunctional enzymes from methanol-grown Pichia pastoris NRRL-Y-7556 and Torulopsis candida Y-11419 and weak precipitin bands with Hansenula polymorpha NRRL-Y-2214. No cross-reaction was observed with isofunctional enzyme derived from methanol-grown Kloeckera sp.  相似文献   

7.
Extracts of Pseudomonas C grown on methanol as 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 Pseudomona 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   

8.
A hollow fiber module was used as a reactor for conversion of ethanol to acetaldehyde in the presence of horse liver alcohol dehydrogenase as catalyst. Mass transport rates for NAD+, the overall acetaldehyde generation rate, catalyst effectiveness factors, and the overall order of the reaction with respect to NAD+ concentration were measured. A coupled-substrate reactor with continuous in situ regeneration of cofactor was also examined. Two substrates of opposite redox state were added simultaneously to the feed stream. NADH and acetaldehyde concentrations were monitored in the effluent stream. The cofactor recycle number, or ratio of moles of product to moles of NADH produced, exceeded 10,000 under certain conditions. While decreasing the NAD+ concentration in the feed stream decreased reactor productivity somewhat, it greatly enhanced the ratio of product formed per mole of NAD+ fed to the reactor. It is suggested that high cofactor costs in dehydrogenase reactors may be overcome with efficient in situ regeneration and secondary recovery and recycling of cofactor from the process stream.  相似文献   

9.
Although the facultatively autotrophic acidophile Thiobacillus acidophilus is unable to grow on formate and formaldehyde in batch cultures, cells from glucose-limited chemostat cultures exhibited substrate-dependent oxygen uptake with these C1-compounds. Oxidation of formate and formaldehyde was uncoupler-sensitive, suggesting that active transport was involved in the metabolism of these compounds. Formate- and formaldehyde-dependent oxygen uptake was strongly inhibited at substrate concentrations above 150 and 400 M, respectively. However, autotrophic formate-limited chemostat cultures were obtained by carefully increasing the formate to glucose ratio in the reservoir medium of mixotrophic chemostat cultures. The molar growth yield on formate (Y=2.5 g ·mol-1 at a dilution rate of 0.05 h-1) and RuBPCase activities in cell-free extracts suggested that T. acidophilus employs the Calvin cycle for carbon assimilation during growth on formate. T. acidophilus was unable to utilize the C1-compounds methanol and methylamine. Formate-dependent oxygen uptake was expressed constitutively under a variety of growth conditions. Cell-free extracts contained both dye-linked and NAD-dependent formate dehydrogenase activities. NAD-dependent oxidation of formaldehyde required reduced glutathione. In addition, cell-free extracts contained a dye-linked formaldehyde dehydrogenase activity. Mixotrophic growth yields were higher than the sum of the heterotrophic and autotrophic yields. A quantitative analysis of the mixotrophic growth studies revealed that formaldehyde was a more effective energy source than formate.  相似文献   

10.
The formaldehyde dehydrogenase (EC 1.2.1.1) from the yeast Pichia pastoris IFP 206 was purified to homogeneity. The enzyme had a molecular weight of 84,000 daltons and was composed of two identical subunits of a molecular weight of 39,000 daltons. The N-terminal end of the subunits is blocked. The protein showed 6,3 free -SH groups per mole and 12,5 in the presence of NAD+. Enzyme stability was increased by addition of glycerol during the purification.

The enzyme activity is NAD+ and glutathione dependent. The reaction product is S formylglutathione. The presence of an S-formylglutathione hydrolase (EC 3.1.2.12) in the cell free extract was detected. The formaldehyde dehydrogenase showed an optimum pH of 7.9 and an optimum temperature of 47°C. The activation energy was 3.2 kcal/mol. The Michaelis constants for NAD+ and S-hydroxymethyl glutathione were respectively 0.24 mM and 0.26 mM.  相似文献   

11.
The formate dehydrogenase from the yeast Pichia pastoris IFP 206 was purified to homogeneity. The protein showed a molecular weight of 68,000 daltons and was composed of two identical subunits. Its amino acid composition was similar to those of other formate dehydrogenases and was characterized by a high content of acidic residues. The N-terminal end of the molecule was probably blocked.

The enzyme activity was NAD+ dependent (NADP+ could not replace NAD+). Its optimum temperature was 47°C and the activation energy 10.8 kcal/mol. The enzyme was active from pH 3.5 to 10.5 with a maximum at pH 7.5. The Michaelis constant for NAD+ and formate were respectively 0.27 and 15mM. The purified enzyme had no S-formylglutathione hydrolase activity, strongly suggesting that the true substrate was formate. NADH, cyanide and azide were strong inhibitors of the enzyme.  相似文献   

12.
Previously, we constructed a glycerol oxidative pathway-deficient mutant strain of Klebsiella pneumoniae by inactivation of glycerol dehydrogenase (dhaD) to eliminate by-product synthesis during production of 1,3-propanediol (1,3-PD) from glycerol. Although by-product formation was successfully blocked in the resultant strain, the yield of 1,3-PD was not enhanced, probably because dhaD disruption resulted in insufficient regeneration of the cofactor NADH essential for the activity of 1,3-PD oxidoreductase (DhaT). To improve cofactor regeneration, in the present study we overexpressed an NAD+-dependent aldehyde dehydrogenase in the recombinant strain. To this end, an aldehyde dehydrogenase AldHk homologous to E. coli AldH but with NAD+-dependent propionaldehyde dehydrogenase activity was identified in K. pneumoniae. Functional analysis revealed that the substrate specificity of AldHk embraced various aldehydes including propionaldehyde, and that NAD+ was preferred over NADP+ as a cofactor. Overexpression of AldHk in the glycerol oxidative pathway-deficient mutant AK/pVOTHk resulted in a 3.6-fold increase (0.57 g l−1 to 2.07 g l−1) in the production of 3-hydroxypropionic acid (3-HP), and a 1.1-fold enhancement (8.43 g l−1 to 9.65 g l−1) of 1,3-PD synthesis, when glycerol was provided as the carbon source, compared to the levels synthesized by the control strain (AK/pVOT). Batch fermentation using AK/pVOTHk showed a significant increase (to 70%, w/w) in conversion of glycerol to the reductive metabolites, 1,3-PD and 3-HP, with no production of by-products except acetate.  相似文献   

13.
The NAD+‐dependent lactate dehydrogenase from Bacillus subtilis (BsLDH) catalyzes the enantioselective reduction of pyruvate to lactate. BsLDH is highly specific to NAD+ and exhibits only a low activity with NADP+ as cofactor. Based on the high activity and good stability of LDHs, these enzymes have been frequently used for the regeneration of NAD+. While an application in the regeneration of NADP+ is not sufficient due to the cofactor preference of the BsLDH. In addition, NADP+‐dependent LDHs have not yet been found in nature. Therefore, a structure‐based approach was performed to predict amino acids involved in the cofactor specificity. Methods of site‐saturation mutagenesis were applied to vary these amino acids, with the aim to alter the cofactor specificity of the BsLDH. Five constructed libraries were screened for improved NADP+ acceptance. The mutant V39R was identified to have increased activity with NADP+ relative to the wild type. V39R was purified and biochemically characterized. V39R showed excellent kinetic properties with NADP(H) and NAD(H), for instance the maximal specific activity with NADPH was enhanced 100‐fold to 90.8 U/mg. Furthermore, a 249‐fold increased catalytic efficiency was observed. Surprisingly, the activity with NADH was also significantly improved. Overall, we were able to successfully apply V39R in the regeneration of NADP+ in an enzyme‐coupled approach combined with the NADP+‐dependent alcohol dehydrogenase from Lactobacillus kefir. We demonstrate for the first time an application of an LDH in the regeneration of NADP+.  相似文献   

14.
The ethanologenic bacterium Zymomonas mobilis ZM4 is of special interest because it has a high ethanol yield. This is made possible by the two alcohol dehydrogenases (ADHs) present in Z. mobilis ZM4 (zmADHs), which shift the equilibrium of the reaction toward the synthesis of ethanol. They are metal-dependent enzymes: zinc for zmADH1 and iron for zmADH2. However, zmADH2 is inactivated by oxygen, thus implicating zmADH2 as the component of the cytosolic respiratory system in Z. mobilis. Here, we show crystal structures of zmADH2 in the form of an apo-enzyme and an NAD+-cofactor complex. The overall folding of the monomeric structure is very similar to those of other functionally related ADHs with structural variations around the probable substrate and NAD+ cofactor binding region. A dimeric structure is formed by the limited interactions between the two subunits with the bound NAD+ at the cleft formed along the domain interface. The catalytic iron ion binds near to the nicotinamide ring of NAD+, which is likely to restrict and locate the ethanol to the active site together with the oxidized Cys residue and several nonpolar bulky residues. The structures of the zmADH2 from the proficient ethanologenic bacterium Z. mobilis, with and without NAD+ cofactor, and modeling ethanol in the active site imply that there is a typical metal-dependent catalytic mechanism.  相似文献   

15.
During glucose‐limited growth, a substantial input of adenosine triphosphate (ATP) is required for the production of β‐lactams by the filamentous fungus Penicillium chrysogenum. Formate dehydrogenase has been confirmed in P. chrysogenum for formate oxidation allowing an extra supply of ATP, and coassimilation of glucose and formate has the potential to increase penicillin production and biomass yield. In this study, the steady‐state metabolite levels and fluxes in response to cofeeding of formate as an auxiliary substrate in glucose‐limited chemostat cultures at the dilution rates (D) of both 0.03 h?1 and 0.05 h?1 are determined to evaluate the quantitative impact on the physiology of a high‐yielding P. chrysogenum strain. It is observed that an equimolar addition of formate is conducive to an increase in both biomass yield and penicillin production at D = 0.03 h?1, while this is not the case at D = 0.05 h?1. In addition, a higher cytosolic redox status (NADH/NAD+), a higher intracellular glucose level, and lower penicillin productivity are only observed upon formate addition at D = 0.05 h?1, which are virtually absent at D = 0.03 h?1. In conclusion, the results demonstrate that the effect of formate as an auxiliary substrate on penicillin productivity in the glucose‐limited chemostat cultivations of P. chrysogenum is not only dependent on the formate/glucose ratio as published before but also on the specific growth rate. The results also imply that the overall process productivity and quality regarding the use of formate should be further explored in an actual industrial‐scale scenario.  相似文献   

16.
Isocitrate dehydrogenase (IDH) activities were measured in mitochondria isolated from aerial parts of 21-day-old spruce (Picea abies L. Karst.) seedlings. Mitochondria were purified by two methods, involving continuous and discontinuous Percoll gradients. Whatever the method of purification, the mitochondrial outer membrane was about 69% intact, and the mitochondria contained very low cytosolic, chloroplastic and peroxisomal contaminations. Nevertheless, as judged by the recovery of fumarase activity, purification on a continuous 28% Percoll gradient gave the best yield in mitochondria, which exhibited a high degree of inner membrane intactness (91%). The purified mitochondria oxidized succinate and malate with good respiratory control and ADP/O ratios. The highest oxidation rate was obtained with succinate as substrate, and malate oxidation was improved (+ 60%) by addition of exogenous NAD+. Experiments using standard respiratory chain inhibitors indicated that, in spruce mitochondria, the alternative pathway was present. Both NAD+-isocitrate dehydrogenase (EC 1.1.1.41) and NADP+-isocitrate dehydrogenase (EC 1.1.1.42) were present in the mitochondrial matrix fraction, and NAD+-IDH activity was about 2-fold higher than NADP+-IDH activity. The NAD+-IDH showed sigmoidal kinetics in response to isocitrate and standard Michaelis-Menten kinetics for NAD+ and Mg2+. The NADP+-IDH, in contrast, displayed lower Km values. For NAD+-IDH the pH optimum was at 7.4, whereas NADP+-IDH exhibited a broad pH optimum between 8.3 and 9. In addition, NAD+-IDH was more thermolabile. Adenine nucleotides and 2-oxoglutarate were found to inhibit NAD(P)+-IDH activities only at high concentrations.  相似文献   

17.
Polyol dehydrogenases of Acetobacter melanogenum were investigated. Three polyol dehydrogenases, i. e. NAD+-linked d-mannitol dehydrogenase, NAD+-linked sorbitol dehydrogenase and NADP+-linked d-mannitol dehydrogenase, in the soluble fraction of the organism were purified 12-fold, 8-fold and 88-fold, respectively, by fractionation with ammonium sulfate and DEAE-cellulose column chromatography. NAD+-linked sorbitol dehydrogenase reduced 5-keto-d-fructose (5KF) to l-sorbose in the presence of NADH, whereas NADP+-linked d-mannitol dehydrogenase reduced the same substrate to d-fructose in the presence of NADPH. It was also shown that NAD+-linked d-mannitol dehydrogenase was specific for the interconversion between d-mannitol and d-fructose and that this enzyme was very unstable in alkaline conditions.  相似文献   

18.
Abstract A partially purified preparation of cyanide hydrolase (cyanidase) from a bacterium, Klebsiella sp., was applied as a biocatalyst in a biosensor system for low-level cyanide detection. In the biosensor system cyanide hydrolase converts cyanide into formate and ammonia. The formate produced in the cyanide degradation was detected with a formate biosensor, in which formate dehydrogenase (FDH; E.C. 1.2.1.2) was co-immobilized with salicylate hydroxylase (SHL; E.C. 1.14.13.1) on a Clark electrode. The principle of the formate sensor is that FDH converts formate into carbon dioxide using -nicotinamide adenine dinucleotide hydrate (NAD+). The corresponding NADH produced is then oxidized to NAD+ by SHL using salicylate and oxygen. The oxygen consumption is monitored with the Clark electrode. The optimum buffer pH and temperature for the enzymatic hydrolysis of potassium cyanide were studied. The preliminary experiments including the pretreatment of cyanide with cyanide hydrolase and then detection by the formate sensor gave a detection limit at 7.3 mol l–1 cyanide. The linear range of the calibration curve was between 30 mol l–1 and 300 mol l–1 cyanide.  相似文献   

19.
Activities of the enzymes of formaldehyde (FA) catabolism in recombinant strains of the methylotrophic yeast Hansenula polymorpha overproducing NAD+- and glutathione-dependent formaldehyde dehydrogenase (FADH) were studied under different cultivation conditions and at elevated FA content. Southern dot-blot analysis confirmed the presence of six to eight copies of the target FLD1 gene in stable recombinant clones of H. polymorpha. Under certain cultivation conditions, the transformants resistant to elevated FA concentrations were shown to produce FADH and other bioanalytically important enzymes: formate dehydrogenase, alcohol dehydrogenase, alcohol oxidase, and formaldehyde reductase. The optimal cultivation conditions for recombinants were determined, resulting in maximum synthesis of FADH: methanol as a carbon source, methylamine as a nitrogen source, FA as an inducer, temperature of 37°C, and cells in the early exponential phase of growth.  相似文献   

20.
This study was carried out to define how the overall rate of reaction would be influenced by different degrees of diffusional resistance to cofactor transport within an oxidoreductase membrane matrix. To accomplish this, 0.7–6.6μM yeast alcohol dehydrogenase was immobilized in an albumin matrix crosslinked with 2.5 or 5.0% glutaraldehyde to give 102–1685 μM thick membranes. The enzyme half-life was at least doubled at pH 7.5 or 8.8 on immobilization. Values of the kinetic constants for the soluble and immobilized enzyme were determined at 25°C and pH 8.8 over the range of 0.01–1.0M bulk solution concentration of ethanol as substrate and 140–1000μM bulk solution concentration of nicotinamide adenine dinucleotide (NAD+) as cofactor, to give essentially single substrate kinetics in NAD+. Equilibrium partitioning of ethanol and NAD+ between the solution and membrane was measured and used in the data analysis. The four kinetic constants for the soluble enzyme agreed with literature values; and all increased with immobilization of the enzyme. The Michaelis constants for NAD+ and for ethanol were greater for the immobilized enzyme. The diffusional resistance to NAD+ transport, presented in terms of the Thiele modulus, showed that the overall rate of reaction was decreased by about 50% even at values of the modulus as low as 2.0.  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