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.
2.
Mannitol biosynthesis in Candida magnoliae HH-01 (KCCM-10252), a yeast strain that is currently used for the industrial production of mannitol, is catalyzed by mannitol dehydrogenase (MDH) (EC 1.1.1.138). In this study, NAD(P)H-dependent MDH was purified to homogeneity from C. magnoliae HH-01 by ion-exchange chromatography, hydrophobic interaction chromatography, and affinity chromatography. The relative molecular masses of C. magnoliae MDH, as determined by sodium dodecyl sulfate-polyacrylamide gel electrophoresis and size-exclusion chromatography, were 35 and 142 kDa, respectively, indicating that the enzyme is a tetramer. This enzyme catalyzed both fructose reduction and mannitol oxidation. The pH and temperature optima for fructose reduction and mannitol oxidation were 7.5 and 37°C and 10.0 and 40°C, respectively. C. magnoliae MDH showed high substrate specificity and high catalytic efficiency (kcat = 823 s−1, Km = 28.0 mM, and kcat/Km = 29.4 mM−1 s−1) for fructose, which may explain the high mannitol production observed in this strain. Initial velocity and product inhibition studies suggest that the reaction proceeds via a sequential ordered Bi Bi mechanism, and C. magnoliae MDH is specific for transferring the 4-pro-S hydrogen of NADPH, which is typical of a short-chain dehydrogenase reductase (SDR). The internal amino acid sequences of C. magnoliae MDH showed a significant homology with SDRs from various sources, indicating that the C. magnoliae MDH is an NAD(P)H-dependent tetrameric SDR. Although MDHs have been purified and characterized from several other sources, C. magnoliae MDH is distinguished from other MDHs by its high substrate specificity and catalytic efficiency for fructose only, which makes C. magnoliae MDH the ideal choice for industrial applications, including enzymatic synthesis of mannitol and salt-tolerant plants.  相似文献   

3.
To improve the fermentation efficiency of Propionibacterium acidipropionici, a semi‐continuous coupled fermentation process was established to achieve co‐production of propionic acid (PA) and succinic acid (SA). First, the optimal proportion of glucose (Glc) and glycerol (Gl) as a mixed carbon source was determined, and the feeding procedure of Gl was optimized to make more energy flow in the direction of product synthesis. Then, ZGD630 anion exchange resin was used for efficient adsorption of PA, thereby eliminating the feedback inhibition effect of PA. Finally, an efficient, coupled fermentation process of P. acidipropionici characterized by membrane separation and chromatography technology was developed. The concentrations of PA and SA reached 62.22 ± 2.32 and 20.45 ± 1.34 g L−1, with corresponding productivity of 0.43 and 0.14 g L−1 h−1, increased by 65.38% and 48.54%, respectively. Membrane separation coupled fermentation of PA and SA could significantly improve the process economics of P. acidipropionici, and has good prospects for industrial application.  相似文献   

4.
Three selective enrichment broths and four selective agar media were evaluated for their ability to support the growth of Pseudomonas pseudomallei both at 35°C and at ambient temperature (range, 20 to 32°C; mean, 25°C). Colony counts of 50 strains of P. pseudomallei and recovery studies with 1 soil strain in 60 simulated soil samples demonstrated that enrichment with Trypticase soy broth incorporating 5 mg of crystal violet per liter and 20 mg of colistin per liter (CVCB) and subculture to Ashdown medium supported the growth of all 50 strains and produced the highest recovery rates with the greatest suppression of other soil flora. An enrichment broth of MacConkey broth (purple) incorporating 10 mg of crystal violet per liter, 5 mg of bromcresol purple per liter, 25 mg of gentamicin per liter, and 650 mg of streptomycin per liter showed greater suppression of soil bacteria than CVCB, but it failed to support the growth of three strains of P. pseudomallei. Recovery rates were essentially the same irrespective of whether the soil samples were incubated at 35°C or at ambient temperature, provided cultures were incubated in protected shade for an extended period. This is an important feature for field work in large-scale epidemiological surveys in which resources are limited.  相似文献   

5.
The recalcitrance of xenobiotics may be caused by an absence of transforming enzymes or by their inability to enter microbial cells. A nondestructive method for differentiating between these two possibilities is described. The solid n-alkanes octadecane (C18) and hexatriacontane (C36) were encapsulated into phosphatidylcholine bilayers (liposomes). The uptake and metabolism rates of encapsulated and unencapsulated substrates were then compared. During 1 h at 25°C, a Pseudomonas isolate took up 1.3% of radiolabeled and unencapsulated C18 (solid state) versus 23.5% of labeled and encapsulated C18. Growth at 25°C occurred with an apparent ks of 2453 ± 148 mg/liter. Liposome encapsulation decreased this Ks to 60 ± 12 mg/liter. At 34°C, growth on C18 (liquid state) occurred with an apparent Ks of 819 ± 83 mg/liter and on the readily available carbon source succinate, Ks values were 80 ± 10 and 13 ± 7 mg/liter at 25 and 34°C, respectively. At 25°C, the isolate grew on C36 with an apparent Ks of 2,698 ± 831 mg/liter. Liposome encapsulation decreased the Ks more than 60-fold to 41 ± 7 mg/liter, resulting in the complete utilization of 400 mg of C36 per liter in 16 h. Since controls excluded the metabolic utilization of phosphatidylcholine, the results clearly identify transport limitation as the cause for C36 recalcitrance.  相似文献   

6.
We isolated thermotolerant Gluconobacter strains that are able to produce 5-keto-d-gluconic acid (5KGA) at 37°C, a temperature at which regular mesophilic 5KGA-producing strains showed much less growth and 5KGA production. The thermotolerant strains produced 2KGA as the major product at both 30 and 37°C. The amount of ketogluconates produced at 37°C was slightly less than the amount produced at 30°C. To improve the yield of 5KGA in these strains, we disrupted flavin adenine dinucleotide-gluconate dehydrogenase (FAD-GADH), which is responsible for 2KGA production. Genes for FAD-GADH were cloned by using inverse PCR and an in vitro cloning strategy. The sequences obtained for three thermotolerant strains were identical and showed high levels of identity to the FAD-GADH sequence reported for the genome of Gluconobacter oxydans 621 H. A kanamycin resistance gene cassette was used to disrupt the FAD-GADH genes in the thermotolerant strains. The mutant strains produced 5KGA exclusively, and the final yields were over 90% at 30°C and 50% at 37°C. We found that the activity of pyrroloquinoline quinone (PQQ)-dependent glycerol dehydrogenase, which is responsible for 5KGA production, increased in response to addition of PQQ and CaCl2 in vitro when cells were grown at 37°C. Addition of 5 mM CaCl2 to the culture medium of the mutant strains increased 5KGA production to the point where over 90% of the initial substrate was converted. The thermotolerant Gluconobacter strains that we isolated in this study provide a promising new option for industrial 5KGA production.Gluconobacter is a genus of acetic acid bacteria that are able to oxidize a broad range of sugars, sugar alcohols, and sugar acids, and large amounts of the corresponding oxidized products accumulate in the culture medium. Such “incomplete” oxidation is carried out by membrane-bound enzymes, whose catalytic sites face the periplasm. These enzymes catalyze the dehydrogenization of d-glucose, d-sorbitol, d-mannitol, glycerol, d-gluconate, and the keto-d-gluconates. All of these enzymes are firmly attached to the cytoplasmic membrane, and the electrons abstracted from the substrates are passed on to ubiquinone and then to terminal ubiquinol oxidases, forming simple respiratory chains which create the membrane potential necessary to produce biological energy for these microorganisms.The oxidation of d-glucose to ketogluconates is known to be catalyzed by a series of enzymes. Pyrroloquinoline quinone (PQQ)-dependent glucose dehydrogenase oxidizes d-glucose to glucono-δ-lactone, and then gluconolactonase converts the glucono-δ-lactone to d-gluconate. The formation of ketogluconates in Gluconobacter strains has been reported to be catalyzed by two types of membrane-bound gluconate dehydrogenases (GADH) (10). One type is flavin adenine dinucleotide (FAD)-GADH, an FAD-containing, 2-keto-d-gluconate (2KGA)-producing enzyme, and the other type is a PQQ-containing, 5-keto-d-gluconate (5KGA)-producing enzyme. The former enzyme has three subunits: an FAD-containing dehydrogenase, a c-type cytochrome subunit containing three hemes, and a small subunit of unknown function (17). The latter enzyme, which produces 5KGA, is identical to the PQQ-containing polyol dehydrogenase (9), which is known as d-arabitol dehydrogenase (1), d-sorbitol dehydrogenase (20), or PQQ-dependent glycerol dehydrogenase (PQQ-GLDH) (2). PQQ-GLDH has broad substrate specificity but high regio- and stereospecificity, and it catalyzes reactions as predicted by the Bertrand-Hudson rule. This enzyme can oxidize d-gluconate only at the C-5 position to produce 5KGA from d-gluconate; however, the affinity of the enzyme for d-gluconate is quite low. The gene encoding this enzyme was cloned from Gluconobacter suboxydans IFO 3255 (11), and two open reading frames (ORFs) were found. One of these ORFs is believed to encode a hydrophobic protein with five membrane-spanning regions, and the other encodes a dehydrogenase subunit similar to that found in several PQQ-dependent enzymes, particularly the PQQ domain of membrane-bound glucose dehydrogenase. In contrast, 2KGA reductase and 5KGA reductase, the NADPH-dependent enzymes located in the cytoplasm, are thought to be involved in gluconate metabolism in the assimilation of 2KGA and 5KGA.5KGA is a useful raw material for the production of tartaric acid and xylaric acid and is used as a precursor for the synthesis of a number of flavor compounds, including 4-hydroxy-5- methyl-2,3-dihydrofuranone-3 (15). Moreover, it has been reported that 5KGA can be used to produce vitamin C by Gray''s method (6, 7), which is different from Reichstein''s method, which is now commonly used in industry. Reichstein''s method requires the use of high temperatures and an organic solvent in processing; however, Gray''s method does not.Most Gluconobacter strains produce both 2KGA and 5KGA from d-gluconate. Thus, production of 5KGA by Gluconobacter species generates 2KGA as a major by-product, and production of the two ketogluconates is competitive in vivo. Recently, an FAD-GADH-defective mutant strain of Gluconobacter oxydans 621 H which produced almost exclusively 5KGA from d-glucose was discovered (5). However, the optimum temperature for production of 5KGA in this mesophilic strain was around 20°C (19). For cost-effective industrial synthesis of 5KGA, we sought to develop a Gluconobacter strain which is able to produce 5KGA at higher temperatures, such as 37°C, in order to reduce the cost of cooling during fermentation.We successfully isolated thermotolerant Gluconobacter strains that are able to produce 5KGA at 37°C. We cloned the FAD-GADH gene and constructed FAD-GADH-defective mutants that produced almost exclusively 5KGA from d-gluconate at both ambient temperatures and higher temperatures up to 37°C. We believe that the thermotolerant strains reported in this study should be useful for industrial 5KGA production.  相似文献   

7.
The Jacob and Monod scheme for the regulation of enzyme formation leads to the following relation between the relative rate of enzyme synthesis α and cellular effector concentration E (the lower sign is for repressible systems): log (α/1 - α - αb) = ± n log [E] + log αb ± log K1. This equation permits linear plotting of experimental data and the evaluation of three quantities: n, the number of effector molecules combining with a repressor molecule, K1, the dissociation constant of this interaction and K2/Rt, the ratio of repressor-operator dissociation constant to total repressor concentration. Measurements on the repression of alkaline phosphatase in Escherichia coli as a function of phosphate concentration are reported and fit the proposed equation with n = 1, indicating that the binding of a single phosphate to the repressor species may be sufficient to cause repression. K1 of this interaction was found to be 0.58 ±0.11 × 10-3 M. The available data regarding the enzymes of the lac operon in a variety of E. coli strains, and several other enzymes are analyzed. It is confirmed that the lac repressor interacts with 2 isopropyl thiogalactoside (IPTG) molecules to relieve repression with a K1 = 50 ±20 × 10-12 M2. In some strains, separate binding constants for the first and second IPTG molecules can be evaluated.  相似文献   

8.
Corynebacterium glutamicum, an established microbial cell factory for the biotechnological production of amino acids, was recently genetically engineered for aerobic succinate production from glucose in minimal medium. In this work, the corresponding strains were transformed with plasmid pVWEx1-glpFKD coding for glycerol utilization genes from Escherichia coli. This plasmid had previously been shown to allow growth of C. glutamicum with glycerol as sole carbon source. The resulting strains were tested in minimal medium for aerobic succinate production from glycerol, which is a by-product in biodiesel synthesis. The best strain BL-1/pVWEx1-glpFKD formed 79 mM (9.3 g l−1) succinate from 375 mM glycerol, representing 42% of the maximal theoretical yield under aerobic conditions. A specific succinate production rate of 1.55 mmol g−1 (cdw) h−1 and a volumetric productivity of 3.59 mM h−1 were obtained, the latter value representing the highest one currently described in literature. The results demonstrate that metabolically engineered strains of C. glutamicum are well suited for aerobic succinate production from glycerol.  相似文献   

9.
Glutamate metabolism is linked to a number of fundamental metabolic pathways such as amino acid metabolism, the TCA cycle, and glutathione (GSH) synthesis. In the yeast Saccharomyces cerevisiae, glutamate is synthesized from α-ketoglutarate by two NADP+-dependent glutamate dehydrogenases (NADP-GDH) encoded by GDH1 and GDH3. Here, we report the relationship between the function of the NADP-GDH and stress-induced apoptosis. Gdh3-null cells showed accelerated chronological aging and hypersusceptibility to thermal and oxidative stress during stationary phase. Upon exposure to oxidative stress, Gdh3-null strains displayed a rapid loss in viability associated with typical apoptotic hallmarks, i.e. reactive oxygen species accumulation, nuclear fragmentation, DNA breakage, and phosphatidylserine translocation. In addition, Gdh3-null cells, but not Gdh1-null cells, had a higher tendency toward GSH depletion and subsequent reactive oxygen species accumulation than did WT cells. GSH depletion was rescued by exogenous GSH or glutamate. The hypersusceptibility of stationary phase Gdh3-null cells to stress-induced apoptosis was suppressed by deletion of GDH2. Promoter swapping and site-directed mutagenesis of GDH1 and GDH3 indicated that the necessity of GDH3 for the resistance to stress-induced apoptosis and chronological aging is due to the stationary phase-specific expression of GDH3 and concurrent degradation of Gdh1 in which the Lys-426 residue plays an essential role.  相似文献   

10.
High-level expression of soluble recombinant human hemoglobin (rHb) in Escherichia coli was obtained with several hemoglobin variants. Under identical conditions, two rHbs containing the Presbyterian mutation (Asn-108→Lys) in β-globin accumulated to approximately twofold less soluble globin than rHbs containing the corresponding wild-type β-globin subunit accumulated. The β-globin Providence(asp) mutation (Lys-82→Asp) significantly improved soluble rHb accumulation compared to the wild-type β-globin subunit and restored soluble accumulation of rHbs containing the Presbyterian mutation to wild-type levels. The Providenceasp substitution introduced a negatively charged residue into the normally cationic 2,3-bisphosphoglycerate binding pocket, potentially reducing the electrostatic repulsion in the absence of the polyanion. The average soluble globin accumulation when there was coexpression of di-α-globin and β-Lys-82→Asp-globin (rHb9.1) and heme was present in at least a threefold molar excess was 36% ± 3% of the soluble cell protein in E. coli. The average total accumulation (soluble globin plus insoluble globin) was 56% ± 7% of the soluble cell protein. Fermentations yielded 6.0 ± 0.3 g of soluble rHb9.1 per liter 16 h after induction and 6.4 ± 0.2 g/liter 24 h after induction. The average total globin yield was 9.4 g/liter 16 h after induction. High-level accumulation of soluble rHb in E. coli depends on culture conditions, the protein sequence, and the molar ratio of the heme cofactor added.  相似文献   

11.
The mechanism of insulin dysregulation in children with hyperinsulinism associated with inactivating mutations of short-chain 3-hydroxyacyl-CoA dehydrogenase (SCHAD) was examined in mice with a knock-out of the hadh gene (hadh−/−). The hadh−/− mice had reduced levels of plasma glucose and elevated plasma insulin levels, similar to children with SCHAD deficiency. hadh−/− mice were hypersensitive to oral amino acid with decrease of glucose level and elevation of insulin. Hypersensitivity to oral amino acid in hadh−/− mice can be explained by abnormal insulin responses to a physiological mixture of amino acids and increased sensitivity to leucine stimulation in isolated perifused islets. Measurement of cytosolic calcium showed normal basal levels and abnormal responses to amino acids in hadh−/− islets. Leucine, glutamine, and alanine are responsible for amino acid hypersensitivity in islets. hadh−/− islets have lower intracellular glutamate and aspartate levels, and this decrease can be prevented by high glucose. hadh−/− islets also have increased [U-14C]glutamine oxidation. In contrast, hadh−/− mice have similar glucose tolerance and insulin sensitivity compared with controls. Perifused hadh−/− islets showed no differences from controls in response to glucose-stimulated insulin secretion, even with addition of either a medium-chain fatty acid (octanoate) or a long-chain fatty acid (palmitate). Pull-down experiments with SCHAD, anti-SCHAD, or anti-GDH antibodies showed protein-protein interactions between SCHAD and GDH. GDH enzyme kinetics of hadh−/− islets showed an increase in GDH affinity for its substrate, α-ketoglutarate. These studies indicate that SCHAD deficiency causes hyperinsulinism by activation of GDH via loss of inhibitory regulation of GDH by SCHAD.  相似文献   

12.
The NAD+-dependent isocitrate dehydrogenase from etiolated pea (Pisum sativum L.) mitochondria was purified more than 200-fold by dye-ligand binding on Matrix Gel Blue A and gel filtration on Superose 6. The enzyme was stabilized during purification by the inclusion of 20% glycerol. In crude matrix extracts, the enzyme activity eluted from Superose 6 with apparent molecular masses of 1400 ± 200, 690 ± 90, and 300 ± 50 kD. During subsequent purification steps the larger molecular mass species disappeared and an additional peak at 94 ± 16 kD was evident. The monomer for the enzyme was tentatively identified at 47 kD by sodium dodecyl-polyacrylamide gel electrophoresis. The NADP+-specific isocitrate dehydrogenase activity from mitochondria eluted from Superose 6 at 80 ± 10 kD. About half of the NAD+ and NADP+-specific enzymes remained bound to the mitochondrial membranes and was not removed by washing. The NAD+-dependent isocitrate dehydrogenase showed sigmodial kinetics in response to isocitrate (S0.5 = 0.3 mm). When the enzyme was aged at 4°C or frozen, the isocitrate response showed less allosterism, but this was partially reversed by the addition of citrate to the reaction medium. The NAD+ isocitrate dehydrogenase showed standard Michaelis-Menten kinetics toward NAD+ (Km = 0.2 mm). NADH was a competitive inhibitor (Ki = 0.2 mm) and, unexpectedly, NADPH was a noncompetitive inhibitor (Ki = 0.3 mm). The regulation by NADPH may provide a mechanism for coordination of pyridine nucleotide pools in the mitochondria.  相似文献   

13.
Excised light-grown leaves and etiolated leaves of Hordeum vulgare L. cv Dyan catabolized applied (±)-[2-14C]abscisic acid ([±]-[2-14C]ABA) to phaseic acid (PA), dihydrophaseic acid (DPA), and 2′-hydroxymethyl ABA (2′-HMABA). Identification of these catabolites was made by microchemical methods and by combined capillary gas chromatographymass spectrometry (GC-MS) following high dose feeds of nonlabeled substrate to leaves. Circular dichroism analysis revealed that 2′-HMABA was derived from the (−) enantiomer of ABA. By selecting tissue samples in which endogenous catabolites were undetectable by gas chromatography, it was possible to identify unequivocally ABA catabolites by GC-MS without the need to employ deuteriated substrate to distinguish the (±)-ABA catabolites from the same endogenous compounds. Refeeding studies were used to confirm the catabolic route. The methyl ester of (±)-[214C]-ABA was hydrolyzed efficiently by light-grown leaves of H. vulgare. Leaf age played a significant role in (±)-ABA catabolism, with younger leaves being less able than their older counterparts to catabolize this compound. The catabolism of (±)-ABA was inhibited markedly in water-stressed Hordeum leaves which was characterized by a decreased incorporation of label into 2′-HMABA, DPA, and conjugates. The specific, mixed function oxidase inhibitor, ancymidol, did not inhibit, dramatically, (±)-ABA catabolism in light-grown leaves of Hordeum whereas the 80s ribosome, translational inhibitor, cycloheximide, inhibited this process markedly. The 70s ribosome translational inhibitors, lincomycin and chloramphenicol, were less effective than cycloheximide in inhibiting (±)-ABA catabolism, implying that cytoplasmic protein synthesis is necessary for the catabolism of (±)-ABA in Hordeum leaves whereas chloroplast protein synthesis plays only a minor role. This further suggests that the enzymes involved in (±)-ABA catabolism in this plant are cytoplasmically synthesized and are `turned-over' rapidly, although the enzyme responsible for glycosylating (±)-ABA itself appeared to be stable.  相似文献   

14.
An NADP-dependent dehydrogenase catalyzing the conversion of l-sorbosone to l-ascorbic acid has been isolated from Phaseolus vulgaris L. and Spinacia oleracea L. and partially purified. It is stable at −20°C for up to 8 months. Molecular masses, as determined by gel filtration, were 21 and 29 kilodaltons for bean and spinach enzymes, respectively. Km for sorbosone were 12 ± 2 and 18 ± 2 millimolar and for NADP+, 0.14 ± 0.05 and 1.2 ± 0.5 millimolar, for bean and spinach, respectively. Lycorine, a purported inhibitor of l-ascorbic acid biosynthesis, had no effect on the reaction.  相似文献   

15.
We have shown previously that Escherichia coli K92 produces two different capsular polymers known as CA (colanic acid) and PA (polysialic acid) in a thermoregulated manner. The complex Rcs phosphorelay is largely related to the regulation of CA synthesis. Through deletion of rscA and rscB genes, we show that the Rcs system is involved in the regulation of both CA and PA synthesis in E. coli K92. Deletion of either rcsA or rcsB genes resulted in decreased expression of cps (CA biosynthesis cluster) at 19°C and 37°C, but only CA production was reduced at 19°C. Concerning PA, both deletions enhanced its synthesis at 37°C, which does not correlate with the reduced kps (PA biosynthesis cluster) expression observed in the rcsB mutant. Under this condition, expression of the nan operon responsible for PA catabolism was greatly reduced. Although RcsA and RcsB acted as negative regulators of PA synthesis at 37°C, their absence did not reestablish PA expression at low temperatures, despite the deletion of rcsB resulting in enhanced kps expression. Finally, our results revealed that RcsB controlled the expression of several genes (dsrA, rfaH, h-ns and slyA) involved in the thermoregulation of CA and PA synthesis, indicating that RcsB is part of a complex regulatory mechanism governing the surface appearance in E. coli.  相似文献   

16.
In recent years, glycerol has become an attractive carbon source for microbial processes, as it accumulates massively as a by-product of biodiesel production, also resulting in a decline of its price. A potential use of glycerol in biotechnology is the synthesis of poly(3-hydroxypropionate) [poly(3HP)], a biopolymer with promising properties which is not synthesized by any known wild-type organism. In this study, the genes for 1,3-propanediol dehydrogenase (dhaT) and aldehyde dehydrogenase (aldD) of Pseudomonas putida KT2442, propionate-coenzyme A (propionate-CoA) transferase (pct) of Clostridium propionicum X2, and polyhydroxyalkanoate (PHA) synthase (phaC1) of Ralstonia eutropha H16 were cloned and expressed in the 1,3-propanediol producer Shimwellia blattae. In a two-step cultivation process, recombinant S. blattae cells accumulated up to 9.8% ± 0.4% (wt/wt [cell dry weight]) poly(3HP) with glycerol as the sole carbon source. Furthermore, the engineered strain tolerated the application of crude glycerol derived from biodiesel production, yielding a cell density of 4.05 g cell dry weight/liter in a 2-liter fed-batch fermentation process.  相似文献   

17.
The kinetics of accumulation (per milliliter of culture) of the α- and β- subunits, associated with chloroplast-localized ammonium inducible nicotinamide adenine dinucleotide phosphate-specific glutamate dehydrogenase (NADP-GDH) isoenzymes, were measured during a 3 hour induction of synchronized daughter cells of Chlorella sorokiniana in 29 millimolar ammonium medium under photoautotrophic conditions. The β-subunit holoenzyme(s) accumulated in a linear manner for 3 hours without an apparent induction lag. A 40 minute induction lag preceded the accumulation of the α-subunit holoenzyme(s). After 120 minutes, the α-subunit ceased accumulating and thereafter remained at a constant level (i.e. steady state between synthesis and degradation). From pulsechase experiments, using 35SO4 and immunochemical procedures, the rate of synthesis of the α-subunit was shown to be greater than the β-subunit during the first 80 minutes of induction. The α- and β-subunits had different rates of degradation during the induction period (t½ = 50 versus 150 minutes, respectively) and during the deinduction period (t½ = 5 versus 13.5 minutes) after removal of ammonium from the culture. During deinduction, total NADP-GDH activity decreased with a half-time of 9 minutes. Cycloheximide completely inhibited the synthesis and degradation of both subunits. A model for regulation of expression of the NADP-GDH gene was proposed.  相似文献   

18.
19.
Essential oils are marginally soluble in water, making it challenging to evenly disperse them in foods and resulting in an increased tendency to bind with food lipids and proteins, resulting in lowered antimicrobial efficacy. In the current study, free and nano-dispersed (ND) thymol were compared in terms of their antimicrobial efficacies against Escherichia coli O157:H7 ATCC 43889 and 43894 and Listeria monocytogenes strains Scott A and 101 in apple cider and 2% reduced-fat milk. Apple cider was adjusted to pHs 5.5 and 3.5, and antimicrobial tests were performed at 0.3-, 0.5-, 0.75-, and 1.0-g/liter thymol concentrations at 35, 32, 25, and 4°C. Overall, 0.5 and 1.0 g/liter thymol in nano-dispersion and along with free thymol were inhibitory and bactericidal, respectively, against bacterial strains under all treatment conditions. At pH 5.5, 0.5 g/liter ND thymol was bacteriostatic against L. monocytogenes and E. coli for up to 48 h. At pH 3.5, L. monocytogenes controls did not survive beyond 12 h but E. coli survived and was inhibited by 0.5 g/liter ND thymol after 12 and 48 h in apple cider. E. coli strains were significantly sensitive to 4°C and pH 3.5 (P < 0.05). When bacteria were tested in 2% reduced-fat milk at 35 or 32°C, ND and free thymol demonstrated inhibition at 4.5 g/liter. Thus, the current technology seems to be promising and novel, enabling thymol-containing nano-dispersions that are not only transparent but also effective against pathogens in food applications, especially in clear beverages.  相似文献   

20.
The kinetics of malate dehydrogenase (MDH) catalyzed oxidation/reduction of L-malate/oxaloacetate is pH-dependent due to the proton generated/taken up during the reaction. Previous kinetic studies on the mitochondrial MDH did not yield a consensus kinetic model that explains both substrate and pH dependency of the initial velocity. In this study, we propose, to our knowledge, a new kinetic mechanism to explain kinetic data acquired over a range of pH and substrate concentrations. Progress curves in the forward and reverse reaction directions were obtained under a variety of reactant concentrations to identify associated kinetic parameters. Experiments were conducted at physiologically relevant ionic strength of 0.17 M, pH ranging between 6.5 and 9.0, and at 25°C. The developed model was built on the prior observation of proton uptake upon binding of NADH to MDH, and that the MDH-catalyzed oxidation of NADH may follow an ordered bi-bi mechanism with NADH/NAD binding to the enzyme first, followed by the binding of oxaloacetate/L-malate. This basic mechanism was expanded to account for additional ionic states to explain the pH dependency of the kinetic behavior, resulting in what we believe to be the first kinetic model explaining both substrate and pH dependency of the reaction velocity.  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