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 109 毫秒
1.
In this work, we describe the design of an immobilized enzyme microreactor (IEMR) for use in transketolase (TK) bioconversion process characterization. The prototype microreactor is based on a 200‐μm ID fused silica capillary for quantitative kinetic analysis. The concept is based on the reversible immobilization of His6‐tagged enzymes via Ni‐NTA linkage to surface derivatized silica. For the initial microreactor design, the mode of operation is a stop‐flow analysis which promotes higher degrees of conversion. Kinetics for the immobilized TK‐catalysed synthesis of L ‐erythrulose from substrates glycolaldehyde (GA) and hydroxypyruvate (HPA) were evaluated based on a Michaelis–Menten model. Results show that the TK kinetic parameters in the IEMR (Vmax(app) = 0.1 ± 0.02 mmol min–1, Km(app) = 26 ± 4 mM) are comparable with those measured in free solution. Furthermore, the kcat for the microreactor of 4.1 × 105 s?1 was close to the value for the bioconversion in free solution. This is attributed to the controlled orientation and monolayer surface coverage of the His6‐immobilized TK. Furthermore, we show quantitative elution of the immobilized TK and the regeneration and reuse of the derivatized capillary over five cycles. The ability to quantify kinetic parameters of engineered enzymes at this scale has benefits for the rapid and parallel evaluation of evolved enzyme libraries for synthetic biology applications and for the generation of kinetic models to aid bioconversion process design and bioreactor selection as a more efficient alternative to previously established microwell‐based systems for TK bioprocess characterization. © 2009 American Institute of Chemical Engineers Biotechnol. Prog., 2010  相似文献   

2.
Laccases catalyse the oxidation of a wide range of substrates by a radical-catalyzed reaction mechanism, with a corresponding reduction of oxygen to water in a four-electron transfer process. Due to that, laccases are considered environmentally friendly enzymes, and lately there has been great interest in their use for the transformation and degradation of phenolic compounds. In this work, enzymatic oxidation of catechol and L-DOPA using commercial laccase from Trametes versicolor was performed, in continuously operated microreactors. The main focus of this investigation was to develop a new process for phenolic compounds oxidation, by application of microreactors. For a residence time of 72 s and an inlet oxygen concentration of 0.271 mmol/dm3, catechol conversion of 41.3% was achieved, while approximately the same conversion of L-DOPA (45.0%) was achieved for an inlet oxygen concentration of 0.544 mmol/dm3. The efficiency of microreactor usage for phenolic compounds oxidation was confirmed by calculating the oxidation rates; in the case of catechol oxidation, oxidation rates were in the range from 76.101 to 703.935 g/dm3/d (18–167 fold higher, compared to the case in a macroreactor). To better describe the proposed process, kinetic parameters of catechol oxidation were estimated, using data collected from experiments performed in a microreactor. The maximum reaction rate estimated in microreactor experiments was two times higher than one estimated using the initial reaction rate method from experiments performed in a cuvette. A mathematical model of the process was developed, and validated, using data from independent experiments.  相似文献   

3.
The kinetic properties of a microsomal gill (Na+,K+)-ATPase from the blue crab Callinectes danae were analyzed using the substrate p-nitrophenylphosphate. The (Na+,K+)-ATPase hydrolyzed PNPP obeying cooperative kinetics (n=1.5) at a rate of V=125.4±7.5 U mg−1 with K0.5=1.2±0.1 mmol l−1; stimulation by potassium (V=121.0±6.1 U mg−1; K0.5=2.1±0.1 mmol l−1) and magnesium ions (V=125.3±6.3 U mg−1; K0.5=1.0±0.1 mmol l−1) was cooperative. Ammonium ions also stimulated the enzyme through site–site interactions (nH=2.7) to a rate of V=126.1±4.8 U mg−1 with K0.5=13.7±0.5 mmol l−1. However, K+-phosphatase activity was not stimulated further by K+ plus NH4+ ions. Sodium ions (KI=36.7±1.7 mmol l−1), ouabain (KI=830.3±42.5 μmol l−1) and orthovanadate (KI=34.0±1.4 nmol l−1) completely inhibited K+-phosphatase activity. The competitive inhibition by ATP (KI=57.2±2.6 μmol l−1) of PNPPase activity suggests that both substrates are hydrolyzed at the same site on the enzyme. These data reveal that the K+-phosphatase activity corresponds strictly to a (Na+,K+)-ATPase in C. danae gill tissue. This is the first known kinetic characterization of K+-phosphatase activity in the portunid crab C. danae and should provide a useful tool for comparative studies.  相似文献   

4.
Multienzyme reaction systems with simultaneous coenzyme regeneration have been investigated in a continuously operated membrane reactor at bench scale. NAD(H) covalently bound to polyethylene glycol with a molecular weight of 104 [PEG-10,000-NAD(H)] was used as coenzyme. It could be retained in the membrane reactor together with the enzymes. L -leucine dehydrogenase (LEUDH) was used as catalyze for the reductive amination of α-ketoisocaproate (2-oxo-4-methylpentanoic acid) to L -leucine. Format dehydrogenease (FDH) was used for the regeneration of NADH. Kinetic experiments were carried out to obtain data which could be used in a kinetic model in order to predict the performance of an enzyme membrane reactor for the continuous production of L -leucine. The kinetic constants Vmax and Km of enzymes are all in the same range regardless of whether native NAD(H) or PEG-10,000-NAD(H) is used as coenzyme. L -leucine was produced continuously out of α-ketoisocaproate for 48 days; a maximal conversion of 99.7% was reached. The space-time yield was 324 mmol/L day (or 42.5 g/L day).  相似文献   

5.
The kinetic properties of a microsomal gill (Na+,K+)-ATPase from the freshwater shrimp, Macrobrachium olfersii, acclimated to 21‰ salinity for 10 days were investigated using the substrate p-nitrophenylphosphate. The enzyme hydrolyzed this substrate obeying cooperative kinetics at a rate of 123.6 ± 4.9 U mg− 1 and K0.5 = 1.31 ± 0.05 mmol L− 1. Stimulation of K+-phosphatase activity by magnesium (Vmax = 125.3 ± 7.5 U mg− 1; K0.5 = 2.09 ± 0.06 mmol L− 1), potassium (Vmax = 134.2 ± 6.7 U mg− 1; K0.5 = 1.33 ± 0.06 mmol L− 1) and ammonium ions (Vmax = 130.1 ± 5.9 U mg− 1; K0.5 = 11.4 ± 0.5 mmol L− 1) was also cooperative. While orthovanadate abolished p-nitrophenylphosphatase activity, ouabain inhibition reached 80% (KI = 304.9 ± 18.3 μmol L− 1). The kinetic parameters estimated differ significantly from those for freshwater-acclimated shrimps, suggesting expression of different isoenzymes during salinity adaptation. Despite the ≈2-fold reduction in K+-phosphatase specific activity, Western blotting analysis revealed similar α-subunit expression in gill tissue from shrimps acclimated to 21‰ salinity or fresh water, although expression of phosphate-hydrolyzing enzymes other than (Na+,K+)-ATPase was stimulated by high salinity acclimation.  相似文献   

6.
Summary A new, sensitive and continuous assay for -glucosidase is described exploiting the different angles of rotation for the substrate maltose and the product glucose. Kinetic experiments revealed a very pronounced product inhibition of -glucosidase fromSaccharomyces carlsbergensis with a Ki of 4.85·10–3 M for glucose.The KM of maltose was found to be 37.8·10–3 M. Taking these values, an integral kinetic curve for the enzymatic hydrolysis of maltose was calculated, which is shown to fit the experimental data.Symbols used k1 (min–1) pseudo first-order rate constant (for enzymatic cleavage) - k2 (min–1) rate constant (for mutarotation reaction) - I, P (mol/1) inhibitor (product) concentration - ki (mmol/1) inhibitor constant - KM (mmol/l) Michaelis constant - [M] 589 30 (degree/m · l/mol) molecular rotation at 30°C and 589 nm - s (mmol/l) substrate concentration - R (mmol/mg · min) reaction rate - Vmax (mmol/mg · min) maximal rate - U (mol/min) activity unit (here at 30°C and pH=6.8) Indices O initial value - max maximal value  相似文献   

7.
Two brown algae, Macrocystis pyrifera and Undaria pinnatifida, were employed to remove Cr(III) from aqueous solutions. Both seaweeds were characterized in terms of alginate yields. The alginate contents were 20 and 30% of the dry weight for M. pyrifera and U. pinnatifida, respectively. Kinetics experiments were carried out at different initial pH values. Cr(III) biosorption was affected by the solution pH. The highest metal uptake was found at pH 4 for both biosorbents. Different models were applied to elucidate the rate‐controlling mechanism: pseudo‐first‐order, pseudo‐second‐order, external mass transfer and intra‐particle diffusion. The application of Langmuir, Freundlich and Dubinin–Radushkevich models to the equilibrium data showed a better fitting to the first model. The maximum Cr(III) sorption capacity (qm) and the affinity coefficient (b) were very similar for both biosorbents: 0.77 mmol/g and 1.20 L/mmol for M. pyrifera and 0.74 mmol/g and 1.06 L/mmol for U. pinnatifida. The free energy of the sorption process was estimated using the Dubinin–Radushkevich isotherm. The values indicate that the processes are chemical sorptions. To evaluate the significance of the ion‐exchange mechanism, the light metals (Ca2+, Na+, Mg2+ and K+) and pH were measured during the experiments.  相似文献   

8.
It has been determined catalytic activity of cholinesterases for several insect species (Apanteles glomeratus L., Coccinella septempunctata L, Rhopalosiphumpadi L. and Pieris brassicae L.) that varies in norm from 57 to 199 mmol/hour per 1 gr. It has been calculated the constants of bimolecular interaction (Kii) for insecticides Aztek, Mavric and Bi‐58 new with cholinesterase of the insects. It occurred to be the most sensitivity to Bi‐58 new is peculiar to this ferment in Coccinella 7‐punctata L. Kii (4.54 ± 0.23) 104; whereas cholinesterase of Rhopalosiphum padi L. is the least sensitive to Aztek ‐ Kii (2.9 ± 0.14) 105. Determination of anticholinesterase action coefficient has revealed the value of this indice 118.9 for the preparation Aztek in the system "Rhopalosiphum padi ‐Coccinella 7‐punctata”;. Supposedly, anticholinesterase activity of Aztek is the base of its mechanism action on above‐mentioned species of insects.  相似文献   

9.
l-Alanine dehydrogenase was found in extracts of the antibiotic producer Streptomyces clavuligerus. The enzyme was induced by ammonia, and the level of induction was dependend on the extracellular concentration. l-Alanine was the only amino acid able to induce alanine dehydrogenase. The enzyme was characterized from a 38-fold purified preparation. Pyruvate (K m =1.1 mM), ammonia (K m =20 mM) and NADH (K m =0.14 mM) were required for the reductive amination, and l-alanine (K m =9.1 mM) and NAD (K m =0.5 mM) for the oxidative deaminating reaction. The aminating reaction was inhibited by alanine, serine and NADPH. Alanine inhibited uncompetitively with respect to NADH (K i =1.6 mM) and noncompetitively with respect to ammonia (K i =2.0 mM) and pyruvate (K i =3.0 mM). In the aminating reaction 3-hydroxypyruvate, glyoxylate and 2-oxobutyrate could partially (6–7%) substitute pyruvate. Alanine dehydrogenase from S. clavuligerus differed with respect to its molecular weight (92000) and its kinetic properties from those described for other microorganisms.Abbreviation Alanine-DH l-alanine:NAD oxidoreductase  相似文献   

10.
Kinetic studies of two glucosylation reactions catalyzed by an amyloglucosidase from Rhizopus sp. leading to the synthesis of vanillin-α/β-D-glucoside from D-glucose and vanillin and curcumin-bis-α-D-glucoside from D-glucose and curcumin were investigated in detail. Initial reaction rates were determined from kinetic runs involving different concentrations of D-glucose and vanillin (5?mM to 0.1?M) or D-glucose and curcumin (5?mM to 0.1?M). Graphical double reciprocal plots showed that the kinetics of the two enzyme catalyzed reactions exhibited Ping-Pong Bi-Bi mechanism where competitive substrate inhibition by vanillin/curcumin led to dead-end amyloglucosidase–vanillin/curcumin complexes at higher concentrations of vanillin/curcumin. An attempt to obtain the best fit of this kinetic model through computer simulation yielded in good approximation, the values of four important kinetic parameters, vanillin-α/β-D-glucoside: kcat=35.0±3.2 10?5M?h?1·mg, Ki=10.5±1.1?mM, KmD-glucose=60.0±6.2?mM, Kmvanillin=50.0±4.8?mM; curcumin-bis-α-D-glucoside: kcat=6.07±0.58 10?5M?h?1·mg, Ki=3.0±0.28?mM, KmD-glucose=10.0±0.9?mM, Kmcurcumin=4.6±0.5?mM.  相似文献   

11.
The kinetic constants for 4-aminobutyrate: 2-oxoglutarate aminotransferase (GABA-trans-aminase) and succinate-semialdehyde: NAD+ oxidoreductase (SSA-DH) have been determined using rat brain homogenate. The Michaelis constants for GABA-T at saturated substrate concentrations were calculated to be Kgaba= 1.5 mM, K2-OG= 0.25 mM, KGLU= 620 μM, and KSSA= 87 μm. The Vmax for the reaction using GABA and 2-oxoglutarate (2-OG) as substrates (forward reaction) was found to be 35.2 μmol/ These results indicate that MOPEG is a measure for changes in central noradrenaline turnover and that drugs affect MOPEG in the brain and spinal cord similarly. Fractional rate constants of MOPEG in the rat brain and spinal cord were estimated with the exponential decline curves produced by treatment with pargyline. Turnover rates of 193 pmol/gh and 167 pmol/g These results indicate that MOPEG is a measure for changes in central noradrenaline turnover and that drugs affect MOPEG in the brain and spinal cord similarly. Fractional rate constants of MOPEG in the rat brain and spinal cord were estimated with the exponential decline curves produced by treatment with pargyline. Turnover rates of 193 pmol/g/h and 167 pmol/g/h in the brain and spinal cord respectively were calculated. The kinetics of GABA-T have been shown to be consistent with a Ping Pong Bi Bi mechanism. Substrate inhibition of the forward reaction, through formation of a dead-end complex, was found to occur with 2-OG (Ki 3.3 mM), whereas GABA was found to be a product inhibitor of the reverse reaction (Ki= 0.6 mM). Using the appropriate Haldane relationship, a Keq of 0.04 for GGBA-T was found, indicating that the reaction was strongly biased towards GABA. For SSA-DH, the Km of SSA was determined (9.1 μM) and the Vmax was 27.5 μmol/ These results indicate that MOPEG is a measure for changes in central noradrenaline turnover and that drugs affect MOPEG in the brain and spinal cord similarly. Fractional rate constants of MOPEG in the rat brain and spinal cord were estimated with the exponential decline curves produced by treatment with pargyline. Turnover rates of 193 pmol/g/h and 167 pmol/g These results indicate that MOPEG is a measure for changes in central noradrenaline turnover and that drugs affect MOPEG in the brain and spinal cord similarly. Fractional rate constants of MOPEG in the rat brain and spinal cord were estimated with the exponential decline curves produced by treatment with pargyline. Turnover rates of 193 pmol/g/h and 167 pmol/g/h in the brain and spinal cord respectively were calculated. h. The effect of di-n-propylacetate (DPA) on both GABA-T and SSA-DH was measured. DPA inhibited SSA-DH competitively with respect to SSA, giving a Ki of 0.5 mM. GABA-T was only slightly inhibited. The Ki of DPA for the forward reaction was 23.2 mM with respect to GABA, which was 40-50 times higher than that for SSA-DH. For the reverse reaction the Ki of DPA was found to be nearly the same (15.2 mM with respect to Glu and 22.9 mM with respect to SSA). These results suggest that GABA accumulation in the brain, after administration of DPA in vivo, is caused by SSA-DH inhibition. Two mechanisms are indicated by the data. (1) The higher level of SSA, which results from inhibition of SSA-DH, initiates the reverse reaction of GABA-T, thus increasing the level of GABA via conversion of SSA. (2) The degradation of GABA is inhibited by SSA, since SSA has a strong inhibitory effect on the forward reaction, as calculated from the present data.  相似文献   

12.
Summary The experiments reported here evaluate the capability of isolated intestinal epithelial cells to accomplish net H+ transport in response to imposed ion gradients. In most cases, the membrane potential was kept constant by means of a K+ plus valinomycin voltage clamp in order to prevent electrical coupling of ion fluxes. Net H+ flux across the cellular membrane was examined at pH 6.0 (the physiological lumenal pH) and at pH 7.4 using methylamine distribution or recordings of changes in media pH. Results from both techniques suggest that the cells have an Na+/H+ exchange system in the plasma membrane that is capable of rapid and sustained changes in intracellular pH in response to an imposed Na+ gradient. The kinetics of the Na+/H+ exchange reaction at pH 6.0 [K t for Na+=57mm,V max=42 mmol H+/liter 3OMG (3-O-methylglucose) space/min] are dramatically different from those at pH 7.4 (K t for Na+=15mm,V max=1.7 mmol H+/liter 3OMG space/min). Experiments involving imposed K+ gradients suggest that these cells have negligible K+/H+ exchange capability. They exhibit limited but measurable H+ conductance. Anion exchange for base equivalents was not detected in experiments performed in media nominally free of bicarbonate.  相似文献   

13.
In this study, enzymatic oxidation of hexanol to hexanal (green note fragrance) using NAD+ dependent commercial alcohol dehydrogenase from S. cerevisiae was conducted in continuously operated tubular microreactors with internal volumes of 6 and 13 μL and in a tubular microreactor with a volume of 2 μL that was equipped with internal micromixers. Flow profiles in microchannel were observed in experiments in which the aqueous phase was stained brilliant blue and the hexane was kept colourless. The effects of enzyme and coenzyme inlet concentrations and flow ratios of the immiscible phases on the conversion of hexanol and the volumetric productivity of hexanal were analyzed. Significant improvement in the conversion of hexanol when compared to the classical macroscale process was obtained for c i,hexanol = 5.5 mmol/L, c i,NAD+ = 0.55 mmol/L, and γ i,ADH = 0.092 g/L. In the 6 μL microreactor 11.78% conversion of hexanol was attained after 72 sec, while in the macroscale process 5.3% conversion of hexanol was reached after 180 sec.  相似文献   

14.
Binding of NAD and NADH to dihydrolipoamide dehydrogenase fromEscherichia coli and from pig heart was measured using the spin-labeled analogsN 6-(2,2,6,6-tetramethylpiperidine-4-yl-1-oxyl)-NAD and -NADH. A decrease in the peak amplitudes of the respective EPR spectra results after adding enzyme to the cofactor analogs. With the bacterial enzyme normal hyperbolic saturation behavior with the NAD analog and one binding site per subunit (K s =0.51 mM) are observed, while the NADH analog reveals a sigmoidal binding characteristic. A high-affinity and a low-affinity site (K s =0.087 and 0.33 mM) are found for binding of the NAD analog to the pig heart enzyme and only one type of binding site is observed for the NADH analog (K s =22 µM).  相似文献   

15.
Introduction – Aurones (aureusidin glycosides) are plant flavonoids that provide yellow colour to the flowers of some ornamental plants. In this study we analyse the capacity of tyrosinase to catalyse the synthesis of aureusidin by tyrosinase from the chalcone THC (2′,4′,6′,4–tetrahydroxychalcone). Objective – To develop a simple continuous spectrophotometric assay for the analysis of the spectrophotometric and kinetic characteristics of THC oxidation by tyrosinase. Methodology – THC oxidation was routinely assayed by measuring the increase in absorbance at 415 nm vs. reaction time. Results – According to the mechanism proposed for tyrosinase, the enzymatic reaction involves the o‐hydroxylation of the monophenol THC to the o‐diphenol (PHC, 2′,4′,6′,3,4 – pentahydroxychalcone), which is then oxidised to the corresponding o‐quinone in a second enzymatic step. This product is highly unstable and thus undergoes a series of fast chemical reactions to produce aureusidin. In these experimental conditions, the optimum pH for THC oxidation is 4.5. The progress curves obtained for THC oxidation showed the appearance of a lag period. The following kinetic parameters were also determined: Km = 0.12 mM, Vm = 13 μM/min, Vm/Km = 0.11/min. Conclusion – This method has made it possible to analyse the spectrophotometric and kinetic characteristics of THC by tyrosinase. This procedure has the advantages of a short analysis time, straightforward measurement techniques and reproducibility. In addition, it also allows the study of tyrosinase inhibitors, such as tropolone. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

16.
The aim of this study was to evaluate a physiologically based pharmacokinetic (PBPK) model for predicting PK profiles in humans based on a model refined in rats and humans in vitro uptake‐transport data using valsartan as a probe substrate. Valsartan is eliminated unchanged, mostly through biliary excretion, both in humans and rats. It was, therefore, chosen as model compound to predict in vivo elimination based on in vitro hepatic uptake‐transport data using a fully mechanistic PBPK model. Plated rat and human hepatocytes, and cell lines overexpressing human OATP1B1 and OATP1B3 were used for in vitro uptake experiments. A mechanistic two‐compartment model was used to derive the active and passive transport parameters, namely uptake Michaelis–Menten parameters (Vmax and Km,u) together with passive diffusion (Pdif). These transport parameters were then used as input in a whole body physiologically based pharmacokinetic (PBPK) model. The uptake rate of valsartan was higher for rat hepatocytes (Km,u=28.4±3.7 μM , Vmax=1320±180 pmol/mg/min, and Pdif =1.21±0.42 μl/mg/min) compared to human hepatocytes (Km,u=44.4±14.6 μM , Vmax=304±85 pmol/mg/min, and Pdif=0.724±0.271 μl/mg/min). OATP1B1 and ‐1B3 parameters were correlated to human hepatocyte data, using experimentally established relative activity factors (RAF). Resulting PBPK simulations were compared for plasma‐ (humans and rats) and bile‐ (rats) concentration–time profiles following iv bolus administration of valsartan. Plasma clearances (CLP) for rats and humans were predicted within twofold relative to predictions based on respective in vitro data. The simulations were extended to simulate the impact of either OATP1B1 or ‐1B3 inhibition on plasma profile. The limited data set indicates that the mechanistic model allowed for accurate evaluation of in vitro transport data; and the resulting hepatic uptake transport kinetic parameters enabled the prediction of in vivo PK profiles and plasma clearances, using PBPK modelling. Moreover, the interspecies difference in elimination rate observed in vivo was correctly reflected in the transport parameters determined in vitro.  相似文献   

17.
Both conventional and genetic engineering techniques can significantly improve the performance of animal cell cultures for the large-scale production of pharmaceutical products. In this paper, the effect of such techniques on cell yield and antibody production of two NS0 cell lines is presented. On the one hand, the effect of fed-batch cultivation using dialysis is compared to cultivation without dialysis. Maximum cell density could be increased by a factor of ~5–7 by dialysis fed-batch cultivation. On the other hand, suppression of apoptosis in the NS0 cell line 6A1 bcl-2 resulted in a prolonged growth phase and a higher viability and maximum cell density in fed-batch cultivation in contrast to the control cell line 6A1 (100)3. These factors resulted in more product formation (by a factor ~2). Finally, the adaptive model-based OLFO controller, developed as a general tool for cell culture fed-batch processes, was able to control the fed-batch and dialysis fed-batch cultivations of both cell lines.Abbreviations A membrane area (dm2) - c Glc,F glucose concentration in nutrient feed (mmol L–1) - c Glc,FD glucose concentration in dialysis feed (mmol L–1) - c Glc,i glucose concentration in inner reactor chamber (mmol L–1) - c Glc,o glucose concentration in outer reactor chamber (dialysis chamber) (mmol L–1) - c Lac,FD lactate concentration in dialysis feed (mmol L–1) - c Lac,i lactate concentration in inner reactor chamber (mmol L–1) - c Lac,o lactate concentration in outer reactor chamber (dialysis chamber) (mmol L–1) - c LS,FD limiting substrate concentration in dialysis feed (mmol L–1) - c LS,i limiting substrate concentration in inner reactor chamber (mmol L–1) - c LS,o limiting substrate concentration in outer reactor chamber (dialysis chamber) (mmol L–1) - c Mab monoclonal antibody concentration (mg L–1) - F D feed rate of dialysis feed (L h–1) - F Glc feed rate of nutrient concentrate feed (L h–1) - K d maximum death constant (h–1) - k d,LS death rate constant for limiting substrate (mmol L–1) - k Glc monod kinetic constant for glucose uptake (mmol L–1) - k Lac monod kinetic constant for lactate uptake (mmol L–1) - k LS monod kinetic constant for limiting substrate uptake (mmol L–1) - K Lys cell lysis constant (h–1) - K S,Glc monod kinetic constant for glucose (mmol L–1) - K S,LS monod kinetic constant for limiting substrate (mmol L–1) - µ cell-specific growth rate (h–1) - µ d cell-specific death rate (h–1) - µ d,min minimum cell-specific death rate (h–1) - µ max maximum cell-specific growth rate (h–1) - P Glc membrane permeation coefficient for glucose (dm h–1) - P Lac membrane permeation coefficient for lactate (dm h–1) - P LS membrane permeation coefficient for limiting substrate (dm h–1) - q Glc cell-specific glucose uptake rate (mmol cell–1 h–1) - q Glc,max maximum cell-specific glucose uptake rate (mmol cell–1 h–1) - q Lac cell-specific lactate uptake/production rate (mmol cell–1 h–1) - q Lac,max maximum cell-specific lactate uptake rate (mmol cell–1 h–1) - q LS cell-specific limiting substrate uptake rate (mmol cell–1 h–1) - q LS,max maximum cell-specific limiting substrate uptake rate (mmol cell –1 h–1) - q Mab cell-specific antibody production rate (mg cell–1 h–1) - q MAb,max maximum cell-specific antibody production rate (mg cell–1 h–1) - t time (h) - V i volume of inner reactor chamber (culture chamber) (L) - V o volume of outer reactor chamber (dialysis chamber) (L) - X t total cell concentration (cells L–1) - X viable cell concentration (cells L–1) - Y Lac/Glc kinetic production constant (stoichiometric ratio of lactate production and glucose uptake) (–)  相似文献   

18.
In growing leaves, lack of isoprene synthase (IspS) is considered responsible for delayed isoprene emission, but competition for dimethylallyl diphosphate (DMADP), the substrate for both isoprene synthesis and prenyltransferase reactions in photosynthetic pigment and phytohormone synthesis, can also play a role. We used a kinetic approach based on post‐illumination isoprene decay and modelling DMADP consumption to estimate in vivo kinetic characteristics of IspS and prenyltransferase reactions, and to determine the share of DMADP use by different processes through leaf development in Populus tremula. Pigment synthesis rate was also estimated from pigment accumulation data and distribution of DMADP use from isoprene emission changes due to alendronate, a selective inhibitor of prenyltransferases. Development of photosynthetic activity and pigment synthesis occurred with the greatest rate in 1‐ to 5‐day‐old leaves when isoprene emission was absent. Isoprene emission commenced on days 5 and 6 and increased simultaneously with slowing down of pigment synthesis. In vivo Michaelis–Menten constant (Km) values obtained were 265 nmol m?2 (20 μm ) for DMADP‐consuming prenyltransferase reactions and 2560 nmol m?2 (190 μm ) for IspS. Thus, despite decelerating pigment synthesis reactions in maturing leaves, isoprene emission in young leaves was limited by both IspS activity and competition for DMADP by prenyltransferase reactions.  相似文献   

19.
A previous three phase fluidized sand bed reactor design was improved by adding a draft tube to improve fluidization and submerged effluent tubes for sand separation. The changes had little influence on the oxygen transfer coefficients(K L a), but greatly reduced the aeration rate required for sand suspension. The resulting 12.5 dm3 reactor was operated with 1 h liquid residence time, 10.2dm3/min aeration rate, and 1.7–2.3 kg sand (0.25–0.35 mm diameter) for the degradation of phenol as sole carbon source. The K La of 0.015 s–1 gave more than adequate oxygen transfer to support rates of 180g phenol/h · m3 and 216 g oxygen/h · m3. The biomass-sand ratios of 20–35 mg volatiles/g gave estimated biomass concentrations of 3–6 g volatiles/dm3. Offline kinetic measurements showed weak inhibition kinetics with constants ofK s=0.2 mg phenol/dm3, K o2=0.5 mg oxygen/dm3 and KinI= 122.5 mg phenol/dm3. Very small biofilm diffusion effects were observed. Dynamic experiments demonstrated rapid response of dissolved oxygen to phenol changes below the inhibition level. Experimentally simulated continuous stagewise operation required three stages, each with 1 h residence time, for complete degradation of 300 mg phenol/dm3 · h.  相似文献   

20.
The kinetic and mechanistic details of the interaction between caldendrin, calmodulin and the B‐domain of AKAP79 were determined using a biosensor‐based approach. Caldendrin was found to compete with calmodulin for binding at AKAP79, indicating overlapping binding sites. Although the AKAP79 affinities were similar for caldendrin (KD = 20 n m ) and calmodulin (KD = 30 n m ), their interaction characteristics were different. The calmodulin interaction was well described by a reversible one‐step model, but was only detected in the presence of Ca2+. Caldendrin interacted with a higher level of complexity, deduced to be an induced fit mechanism with a slow relaxation back to the initial encounter complex. It interacted with AKAP79 also in the absence of Ca2+, but with different kinetic rate constants. The data are consistent with a similar initial Ca2+‐dependent binding step for the two proteins. For caldendrin, a second Ca2+‐independent rearrangement step follows, resulting in a stable complex. The study shows the importance of establishing the mechanism and kinetics of protein–protein interactions and that minor differences in the interaction of two homologous proteins can have major implications in their functional characteristics. These results are important for the further elucidation of the roles of caldendrin and calmodulin in synaptic function. Copyright © 2012 John Wiley & Sons, Ltd.  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