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 46 毫秒
1.
Sixteen Tn916-induced mutants of Clostridium acetobutylicum were selected that were defective in the production of acetone and butanol. Formation of ethanol, however, was only partially affected. The strains differed with respect to the degree of solvent formation ability and could be assigned to three different groups. Type I mutants (2 strains) were completely defective in acetone and butanol production and contained one or three copies of Tn916 in the chromosome. Analysis of the mutants for enzymes responsible for solvent production revealed the presence of a formerly unknown, specific acetaldehyde dehydrogenase. The data obtained also strongly indicate that the NADP+-dependent alcohol dehydrogenase is in vivo reponsible for ethanol formation, whereas the NAD+-dependent alcohol dehydrogenase is probably involved in butanol production. No activity of this enzyme together with all other enzymes in the acetone and butanol pathway could be found in type I strains. All tetracycline-resistant mutants obtained did no longer sporulate.Non-standard abbreviations AADC acetoacetate decarboxylase - AcaDH acetaldehyde dehydrogenase - BuaDH butyraldehyde dehydrogenase - CoA-TF acetoacetyl coenzyme A: acetate/butyrate: coenzyme A transferase - NAD-ADH, NAD+ dependent alcohol dehydrogenase - NADP-ADH, NADP+ dependent alcohol dehydrogenase  相似文献   

2.
Structural analysis of glucose dehydrogenase from Haloferax mediterranei revealed that the adenosine 2′-phosphate of NADP+ was stabilized by the side chains of Arg207 and Arg208. To investigate the structural determinants for coenzyme specificity, several mutants involving residues Gly206, Arg207 and Arg208 were engineered and kinetically characterized. The single mutants G206D and R207I were less efficient with NADP+ than the wild type, and the double and triple mutants G206D/R207I and G206D/R207I/R208N showed no activity with NADP+.In the single mutant G206D, the relation kcat/KNAD+ was 1.6 times higher than in the wild type, resulting in an enzyme that preferred NAD+ over NADP+. The single mutation was sufficient to modify coenzyme specificity, whereas other dehydrogenases usually required more than one or two mutations to change coenzyme specificity. However, the highest reaction rates were reached with the double mutant G206D/R207I and with coenzyme NAD+, where the kcat was 1.6 times higher than the kcat of the wild-type enzyme with NADP+. However, catalytic efficiency with NAD+ was lower, as the Km value for coenzyme was 77 times higher than the wild type with NADP+.  相似文献   

3.
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   

4.
In Euglena gracilis, pyruvate:NADP+ oxidoreductase, in addition to the pyruvate dehydrogenase complex, functions for the oxidative decarboxylation of pyruvate in the mitochondria. Furthermore, the 2-oxoglutarate dehydrogenase complex is absent, and instead 2-oxoglutarate decarboxylase is found in the mitochondria. To elucidate the central carbon and energy metabolisms in Euglena under aerobic and anaerobic conditions, physiological significances of these enzymes involved in 2-oxoacid metabolism were examined by gene silencing experiments. The pyruvate dehydrogenase complex was indispensable for aerobic cell growth in a glucose medium, although its activity was less than 1% of that of pyruvate:NADP+ oxidoreductase. In contrast, pyruvate:NADP+ oxidoreductase was only involved in the anaerobic energy metabolism (wax ester fermentation). Aerobic cell growth was almost completely suppressed when the 2-oxoglutarate decarboxylase gene was silenced, suggesting that the tricarboxylic acid cycle is modified in Euglena and 2-oxoglutarate decarboxylase takes the place of the 2-oxoglutarate dehydrogenase complex in the aerobic respiratory metabolism.  相似文献   

5.
Methanobacterium thermoautotrophicum (strain Marburg) was found to contain two malate dehydrogenases, which were partially purified and characterized. One was specific for NAD+ and catalyzed the dehydrogenation of malate at approximately one-third of the rate of oxalacetate reduction, and the other could equally well use NAD+ and NADP+ as coenzyme and catalyzed essentially only the reduction of oxalacetate. Via the N-terminal amino acid sequences, the encoding genes were identified in the genome of M. thermoautotrophicum (strain ΔH). Comparison of the deduced amino acid sequences revealed that the two malate dehydrogenases are phylogenetically only distantly related. The NAD+-specific malate dehydrogenase showed high sequence similarity to l-malate dehydrogenase from Methanothermus fervidus, and the NAD(P)+-using malate dehyrogenase showed high sequence similarity to l-lactate dehydrogenase from Thermotoga maritima and l-malate dehydrogenase from Bacillus subtilis. A function of the two malate dehydrogenases in NADPH:NAD+ transhydrogenation is discussed. Received: 29 December 1997 / Accepted: 4 March 1998  相似文献   

6.
The l-alanine dehydrogenase (ADH) of Anabaena cylindrica has been purified 700-fold. It has a molecular weight of approximately 270000, has 6 sub-units, each of molecular weight approximately 43000, and shows activity both in the aminating and deaminating directions. The enzyme is NADH/NAD+ specific and oxaloacetate can partially substitute for pyruvate. The K m app for NAD+ is 14 M and 60 M at low and high NAD+ concentrations, respectively. The K m app for l-alanine is 0.4 mM, that for pyruvate is 0.11 mM, and that for oxaloacetate is 3.0 mM. The K m app for NH 4 + varies from 8–133 mM depending on the pH, being lowest at high pH levels (pH 8.7 or above). Alanine, serine and glycine inhibit ADH activity in the aminating direction. The enzyme is active both in heterocysts and vegetative cells and activity is higher in nitrogen-starved cultures than in N2-fixing cultures. The data suggest that although alanine is formed by the aminating activity of ADH, entry of newly fixed ammonia into organic combination does not occur primarily via ADH in N2-fixing cultures of A. cylindrica. Ammonia assimilation via ADH may be important in cultures with an excess of available nitrogen. The deaminating activity of the enzyme may be important under conditions of nitrogen-deficiency.Abbreviations ADH alanine dehydrogenase - DEAE diethylamino ethyl cellulose - EDTA ethylenediamine tetraacetic acid - GDH glutamic dehydrogenase - GS glutamine synthetase - GOT aspartate-glutamate aminotransferase - NAD+ nicotinamide adenine dinucleotide - NADH reduced nicotinamide adenine dinucleotide - NADP+ nicotinamide adenine dinucleotide phosphate - NADPH reduced nicotinamide adenine dinucleotide phosphate - SDS sodium dodecyl sulphate - Tris tris(hydroxymethyl) aminomethane  相似文献   

7.
NADP-Utilizing Enzymes in the Matrix of Plant Mitochondria   总被引:9,自引:4,他引:5       下载免费PDF全文
Purified potato tuber (Solanum tuberosum L. cv Bintie) mitochondria contain soluble, highly latent NAD+- and NADP+-isocitrate dehydrogenases, NAD+- and NADP+-malate dehydrogenases, as well as an NADPH-specific glutathione reductase (160, 25, 7200, 160, and 16 nanomoles NAD(P)H per minute and milligram protein, respectively). The two isocitrate dehydrogenase activities, but not the two malate dehydrogenase activities, could be separated by ammonium sulfate precipitation. Thus, the NADP+-isocitrate dehydrogenase activity is due to a separate matrix enzyme, whereas the NADP+-malate dehydrogenase activity is probably due to unspecificity of the NAD+-malate dehydrogenase. NADP+-specific isocitrate dehydrogenase had much lower Kms for NADP+ and isocitrate (5.1 and 10.7 micromolar, respectively) than the NAD+-specific enzyme (101 micromolar for NAD+ and 184 micromolar for isocitrate). A broad activity optimum at pH 7.4 to 9.0 was found for the NADP+-specific isocitrate dehydrogenase whereas the NAD+-specific enzyme had a sharp optimum at pH 7.8. Externally added NADP+ stimulated both isocitrate and malate oxidation by intact mitochondria under conditions where external NADPH oxidation was inhibited. This shows that (a) NADP+ is taken up by the mitochondria across the inner membrane and into the matrix, and (b) NADP+-reducing activities of malate dehydrogenase and the NADP+-specific isocitrate dehydrogenase in the matrix can contribute to electron transport in intact plant mitochondria. The physiological relevance of mitochondrial NADP(H) and soluble NADP(H)-consuming enzymes is discussed in relation to other known mitochondrial NADP(H)-utilizing enzymes.  相似文献   

8.
A halophilic NAD+-dependent 2-aminobutyrate dehydrogenase (EC1.4.1.1) was purified to homogeneity from a crude extract of an extreme halophile, Halobacterium saccharovorum DSM 1137, with a 30% yield. The enzyme had a molecular mass of about 160 kDa and consisted of four identical subunits. It retained more than 70% of the activity after heating at 60 °C for 1 h and kept it at 30 °C for 8 months in the presence of 2 M NaCl. The enzyme showed maximum activity in the presence of 2 M RbCl or KCl. The enzyme required NAD+ as a coenzyme and used -2-aminobutyrate, -alanine, and -norvaline as substrates. The best substrate was -2-aminobutyrate. The optimum pH was 9.3 for the oxidative deamination of -2-aminobutyrate and 8.6 for the reductive amination of 2-ketobutyrate. The Michaelis constants were 1.2 mM for -2-aminobutyrate, 0.16 mM for NAD+, 0.012 mM for NADH, 0.78 mM for 2-ketobutyrate, and 500 mM for ammonia in the presence of 2 M KCl. The Km values for the substrates depended on the concentration of KCl, and the Km values decreased under high salt conditions.  相似文献   

9.
Metabolic engineering of Saccharomyces cerevisiae for xylose fermentation has often relied on insertion of a heterologous pathway consisting of nicotinamide adenine dinucleotide (phosphate) NAD(P)H-dependent xylose reductase (XR) and NAD+-dependent xylitol dehydrogenase (XDH). Low ethanol yield, formation of xylitol and other fermentation by-products are seen for many of the S. cerevisiae strains constructed in this way. This has been ascribed to incomplete coenzyme recycling in the steps catalyzed by XR and XDH. Despite various protein-engineering efforts to alter the coenzyme specificity of XR and XDH individually, a pair of enzymes displaying matched utilization of NAD(H) and NADP(H) was not previously reported. We have introduced multiple site-directed mutations in the coenzyme-binding pocket of Galactocandida mastotermitis XDH to enable activity with NADP+, which is lacking in the wild-type enzyme. We describe four enzyme variants showing activity for xylitol oxidation by NADP+ and NAD+. One of the XDH variants utilized NADP+ about 4 times more efficiently than NAD+. This is close to the preference for NADPH compared with NADH in mutants of Candida tenuis XR. Compared to an S. cerevisiae-reference strain expressing the genes for the wild-type enzymes, the strains comprising the gene encoding the mutated XDH in combination a matched XR mutant gene showed up to 50% decreased glycerol yield without increase in ethanol during xylose fermentation.  相似文献   

10.
The hyperthermophilic archaeon Pyrococcus furiosus was grown on pyruvate as carbon and energy source. The enzymes involved in gluconeogenesis were investigated. The following findings indicate that glucose-6-phosphate formation from pyruvate involves phosphoenolpyruvate synthetase, enzymes of the Embden-Meyerhof pathway and fructose-1,6-bisphosphate phosphatase.Cell extracts of pyruvate-grown P.furiosus contained the following enzyme activities: phosphoenolpyruvate synthetase (0.025 U/mg, 50 °C), enolase (0.9 U/mg, 80 °C), phosphoglycerate mutase (0.13 U/mg, 55 °C), phosphoglycerate kinase (0.01 U/mg, 50 °C), glyceraldehyde-3-phosphate dehydrogenase reducing either NADP+ or NAD+ (NADP+: 0.019 U/mg, NAD+: 0.009 U/mg; 50 °C), triosephosphate isomerase (1.4 U/mg, 50 °C), fructose-1,6-bisphosphate aldolase (0.0045 U/mg, 55 °C), fructose-1,6-bisphosphate phosphatase (0.026 U/mg, 75 °C), and glucose-6-phosphate isomerase (0.22 U/mg, 50 °C). Kinetic properties (V max values and apparent K m values) of the enzymes indicate that they operate in the direction of sugar synthesis. The specific enzyme activities of phosphoglycerate kinase, glyceraldehyde-3-phosphate dehydrogenase (NADP+-reducing) and fructose-1,6-bisphosphate phosphatase in pyruvate-grown P. furiosus were by a factor of 3, 10 and 4, respectively, higher as compared to maltose-grown cells suggesting that these enzymes are induced under conditions of gluconeogenesis. Furthermore, cell extracts contained ferredoxin: NADP+ oxidoreductase (0.023 U/mg, 60 °C); phosphoenolpyruvate carboxylase (0.018 U/mg, 50 °C) acts as an anaplerotic enzyme.Thus, in P. furiosus sugar formation from pyruvate involves reactions of the Embden-Meyerhof pathway, whereas sugar degradation to pyruvate proceeds via a modified non-phosphorylated Entner-Doudoroff pathway.  相似文献   

11.
DEAE-cellulose chromatography of extracts of free-living Rhizobium meliloti cells revealed separate NAD+-dependent and NADP+-dependent malic enzyme activities. The NAD+ malic enzyme exhibited more activity with NAD+ as cofactor, but also showed some activity with NADP+. The NADP+ malic enzyme only showed activity when NADP+ was supplied as cofactor. Three independent transposon-induced mutants of R. meliloti which lacked NADP+ malic enzyme activity (dme) but retained NADP+ malic enzyme activity were isolated. In an otherwise wild-type background, the dme mutations did not alter the carbon utilization phenotype; however, nodules induced by these mutants failed to fix N2. Structurally, these nodules appeared to develop like wild-type nodules up to the stage where N2-fixation would normally begin. These results support the proposal that NAD+ malic enzyme, together with pyruvate dehydrogenase, functions in the generation of acetyl-CoA required for TCA cycle function in N2-fixing bacteroids which metabolize C4-dicarboxylic acids supplied by the plant.  相似文献   

12.
Alanine dehydrogenase was purified to near homogeneity from cell-free extract of Streptomyces aureofaciens, which produces tetracycline. The molecular weight of the enzyme determined by size-exclusion high-performance liquid chromatography was 395 000. The molecular weight determined by sodium dodecyl sulfate gel electrophoresis was 48 000, indicating that the enzyme consists of eight subunits with similar molecular weight. The isoelectric point of alanine dehydrogenase is 6.7. The pH optimum is 10.0 for oxidative deamination of L-alanine and 8.5 for reductive amination of pyruvate. K M values were 5.0 mM for L-alanine and 0.11 mM for NAD+. K M values for reductive amination were 0.56 mM for pyruvate, 0.029 mM for NADH and 6.67 mM for NH4Cl.Abbreviation AlaDH alanine dehydrogenase  相似文献   

13.
A recombinant Saccharomyces cerevisiae strain transformed with xylose reductase (XR) and xylitol dehydrogenase (XDH) genes from Pichia stipitis has the ability to convert xylose to ethanol together with the unfavorable excretion of xylitol, which may be due to cofactor imbalance between NADPH-preferring XR and NAD+-dependent XDH. To reduce xylitol formation, we have already generated several XDH mutants with a reversal of coenzyme specificity toward NADP+. In this study, we constructed a set of recombinant S. cerevisiae strains with xylose-fermenting ability, including protein-engineered NADP+-dependent XDH-expressing strains. The most positive effect on xylose-to-ethanol fermentation was found by using a strain named MA-N5, constructed by chromosomal integration of the gene for NADP+-dependent XDH along with XR and endogenous xylulokinase genes. The MA-N5 strain had an increase in ethanol production and decrease in xylitol excretion compared with the reference strain expressing wild-type XDH when fermenting not only xylose but also mixed sugars containing glucose and xylose. Furthermore, the MA-N5 strain produced ethanol with a high yield of 0.49 g of ethanol/g of total consumed sugars in the nonsulfuric acid hydrolysate of wood chips. The results demonstrate that glucose and xylose present in the lignocellulosic hydrolysate can be efficiently fermented by this redox-engineered strain.  相似文献   

14.
Two distinct dihydrolipoamide dehydrogenases (E3s, EC 1.8.1.4) have been detected in pea (Pisum sativum L. cv. Little Marvel) leaf extracts and purified to at or near homogeneity. The major enzyme, a homodimer with an apparent subunit Mr value 56 000 (80–90% of overall activity), corresponded to the mitochondrial isoform studied previously, as confirmed by electrospray mass spectrometry and N-terminal sequence analysis. The minor activity (10–20%), which also behaved as a homodimer, copurified with chloroplasts, and displayed a lower subunit Mr value of 52 000 which was close to the Mr value of 52 614±9.89 Da determined by electrospray mass spectrometry. The plastidic enzyme was also present at low levels in root extracts where it represented only 1–2% of total E3 activity. The specific activity of the chloroplast enzyme was three-to fourfold lower than its mitochondrial counterpart. In addition, it displayed a markedly higher affinity for NAD+ and was more sensitive to product inhibition by NADH. It exhibited no activity with NADP+ as cofactor nor was it inhibited by the presence of high concentrations of NADP+ or NADPH. Antibodies to the mitochondrial enzyme displayed little or no cross-reactivity with its plastidic counterpart and available amino acid sequence data were also suggestive of only limited sequence similarity between the two enzymes. In view of the dual location of the pyruvate dehydrogenase multienzyme complex (PDC) in plant mitochondria and chloroplasts, it is likely that the distinct chloroplastic E3 is an integral component of plastidic PDC, thus representing the first component of this complex to be isolated and characterised to date.Abbreviations E1 pyruvate dehydrogenase - E2 dihydrolipoamide acetyltransferase - E3 dihydrolipoamide dehydrogenase - PDC pyruvate dehydrogenase complex - OGDC 2-oxoglutarate dehydrogenase complex - GDC glycine decarboxylase complex - SDS-PAGE sodium dodecyl sulphate/polyacrylamide gel electrophoresis - TDP thiamine diphosphate - Mr relative molecular mass J.G.L. is grateful to the Biotechnology and Biological Sciences Research Council (BBSRC), U.K. for continuing financial support. M.C. is the holder of a BBSRC-funded earmarked Ph.D. studentship.  相似文献   

15.
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   

16.
Thermoplasma acidophilum utilizes l-rhamnose as a sole carbon source. To determine the metabolic pathway of l-rhamnose in Archaea, we identified and characterized l-rhamnose dehydrogenase (RhaD) in T. acidophilum. Ta0747P gene, which encodes the putative T. acidophilum RhaD (Ta_RhaD) enzyme belonging to the short-chain dehydrogenase/reductase family, was expressed in E. coli as an active enzyme catalyzing the oxidation of l-rhamnose to l-rhamnono-1,4-lactone. Analysis of catalytic properties revealed that Ta_RhaD oxidized l-rhamnose, l-lyxose, and l-mannose using only NADP+ as a cofactor, which is different from NAD+/NADP+-specific bacterial RhaDs and NAD+-specific eukaryal RhaDs. Ta_RhaD showed the highest activity toward l-rhamnose at 60 °C and pH 7. The K m and k cat values were 0.46 mM, 1,341.3 min−1 for l-rhamnose and 0.1 mM, 1,027.2 min−1 for NADP+, respectively. Phylogenetic analysis indicated that branched lineages of archaeal RhaD are quite distinct from those of Bacteria and Eukarya. This is the first report on the identification and characterization of NADP+-specific RhaD.  相似文献   

17.
The gene encoding isocitrate dehydrogenase (IDH) of Methylococcus capsulatus (McIDH) was cloned and overexpressed in Escherichia coli. The purified enzyme was NAD+-dependent with a thermal optimum for activity at 55–60°C and an apparent midpoint melting temperature (T m) of 70°C. Analytical ultracentrifugation (AUC) revealed a homotetrameric state, and McIDH thus represents the first homotetrameric NAD+-dependent IDH that has been characterized. Based on a structural alignment of McIDH and homotetrameric homoisocitrate dehydrogenase (HDH) from Thermus thermophilus (TtHDH), we identified the clasp-like domain of McIDH as a likely site for tetramerization. McIDH showed moreover, higher sequence identity (48%) to TtHDH than to previously characterized IDHs. Putative NAD+-IDHs with high sequence identity (48–57%) to McIDH were however identified in a variety of bacteria showing that NAD+-dependent IDHs are indeed widespread within the domain, Bacteria. Phylogenetic analysis including these new sequences revealed a close relationship with eukaryal allosterically regulated NAD+-IDH and the subfamily III of IDH was redefined to include bacterial NAD+- and NADP+-dependent IDHs. This apparent relationship suggests that the mitochondrial genes encoding NAD+-IDH are derived from the McIDH-like IDHs.  相似文献   

18.
A sensitive isotope exchange method was developed to assess the requirements for and compartmentation of pyruvate and oxalacetate production from malate in proliferating and nonproliferating human fibroblasts. Malatedependent pyruvate production (malic enzyme activity) in the particulate fraction containing the mitochondria was dependent on either NAD+ or NADP+. The production of pyruvate from malate in the soluble, cytosolic fraction was strictly dependent on NADP+. Oxalacetate production from malate (malate dehydrogenase, EC 1.1.1.37) in both the particulate and soluble fraction was strictly dependent on NAD+. Relative to nonproliferating cells, NAD+-linked malic enzyme activity was slightly reduced and the NADP+-linked activity was unchanged in the particulate fraction of serum-stimulated, exponentially proliferating cells. However, a reduced activity of particulate malate dehydrogenase resulted in a two-fold increase in the ratio of NAD(P)+-linked malic enzyme to NAD+-linked malate dehydrogenase activity in the particulate fraction of proliferating fibroblasts. An increase in soluble NADP+-dependent malic enzyme activity and a decrease in NAD+-linked malate dehydrogenase indictated an increase in the ratio of pyruvate-producing to oxalacetate-producing malate oxidase activity in the cytosol of proliterating cells. These coordinate changes may affect the relative amount of malate that is oxidized to oxalacetate and pyruvate in proliferating cells and, therefore, the efficient utilization of glutamine as a respiratory fuel during cell proliferation.  相似文献   

19.
1. Aerobically grown yeast having a high activity of glyoxylate-cycle, citric acid-cycle and electron-transport enzymes was transferred to a medium containing 10% glucose. After a lag phase of 30min. the yeast grew exponentially with a mean generation time of 94min. 2. The enzymes malate dehydrogenase, isocitrate lyase, succinate–cytochrome c oxidoreductase and NADH–cytochrome c oxidoreductase lost 45%, 17%, 27% and 46% of their activity respectively during the lag phase. 3. When growth commenced pyruvate kinase, pyruvate decarboxylase, alcohol dehydrogenase, glutamate dehydrogenase (NADP+-linked) and NADPH–cytochrome c oxidoreductase increased in activity, whereas aconitase, isocitrate dehydrogenase (NAD+- and NADP+-linked), α-oxoglutarate dehydrogenase, fumarase, malate dehydrogenase, succinate–cytochrome c oxidoreductase, NADH–cytochrome c oxidoreductase, NADH oxidase, NADPH oxidase, cytochrome c oxidase, glutamate dehydrogenase (NAD+-linked), glutamate–oxaloacetate transaminase, isocitrate lyase and glucose 6-phosphate dehydrogenase decreased. 4. During the early stages of growth the loss of activity of aconitase, α-oxoglutarate dehydrogenase, fumarase and glucose 6-phosphate dehydrogenase could be accounted for by dilution by cell division. The lower rate of loss of activity of isocitrate dehydrogenase (NAD+- and NADP+-linked), glutamate dehydrogenase (NAD+-linked), glutamate–oxaloacetate transaminase, NADPH oxidase and cytochrome c oxidase implies their continued synthesis, whereas the higher rate of loss of activity of malate dehydrogenase, isocitrate lyase, succinate–cytochrome c oxidoreductase, NADH–cytochrome c oxidoreductase and NADH oxidase means that these enzymes were actively removed. 5. The mechanisms of selective removal of enzyme activity and the control of the residual metabolic pathways are discussed.  相似文献   

20.
An NAD+-dependent xylitol dehydrogenase (XDH) from Fusarium oxysporum, a key enzyme in the conversion of xylose to ethanol, was purified to homogeneity and characterised. It was homodimeric with a subunit of M r 48 000, and pI 3.6. It was optimally active at 45 °C and pH 9–10. It was fully stable at pH 6–7 for 24 h and 30 °C. K m values for d-xylitol and NAD+ were 94 mM and 0.14 mM, respectively. Mn2+ at 10 mM increased XDH activity 2-fold and Cu2+ at 10 mM inhibited activity completely.  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