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 375 毫秒
1.
A stable immobilized preparation of alcohol dehydrogenase (ADH) (EC 1.1.1.1) was obtained by entrapment of ADH-containing Saccharomyces cerevisiae cells in polyacrylamide, polymerized by gamma-rays (100 kR). The permeability barrier for the substrate through the cell membrane was found to be eliminated on entrapment. The stability characteristics, pH-activity profile and other properties of the entrapped ADH are presented. A four-fold enhancement in Km for NAD+ was observed on entrapment, whereas Km for ethanol was not altered.  相似文献   

2.
The influences of nonuniform activity distribution within a porous solid support on the apparent kinetic parameters, Vmapp and Kmapp, of immobilized enzyme reactions following the Michaelis-Menten kinetics were theoretically investigated. As the enzyme is distributed to the neighborhood of the external surface of the support, Vmapp and Kmapp approach their respective intrinsic values over a wide range of substrate concentration. There is a close relationship between the nonuniform distribution and internal diffusional resistance. Changes in these two factors provide similar effects on Vmapp and Kmapp. As long as the immobilized enzyme reaction follows Michaelis-Menten kinetics, the nonuniform activity distribution never makes Kmapp less than its intrinsic value.  相似文献   

3.
1H NMR spectroscopy was used to follow the cleavage of sucrose by invertase. The parameters of the enzyme's kinetics, Km and Vmax, were directly determined from progress curves at only one concentration of the substrate. For comparison with the classical Michaelis-Menten analysis, the reaction progress was also monitored at various initial concentrations of 3.5 to 41.8 mM. Using the Lambert W function the parameters Km and Vmax were fitted to obtain the experimental progress curve and resulted in Km = 28 mM and Vmax = 13 μM/s. The result is almost identical to an initial rate analysis that, however, costs much more time and experimental effort. The effect of product inhibition was also investigated. Furthermore, we analyzed a much more complex reaction, the conversion of farnesyl diphosphate into (+)-germacrene D by the enzyme germacrene D synthase, yielding Km = 379 μM and kcat = 0.04 s− 1. The reaction involves an amphiphilic substrate forming micelles and a water insoluble product; using proper controls, the conversion can well be analyzed by the progress curve approach using the Lambert W function.  相似文献   

4.
The diffusion coefficient (D m) and material/air partition coefficient (K) are two key parameters characterizing the formaldehyde and volatile organic compounds (VOC) sorption behavior in building materials. By virtue of the sorption process in airtight chamber, this paper proposes a novel method to measure the two key parameters, as well as the convective mass transfer coefficient (h m). Compared to traditional methods, it has the following merits: (1) the K, D m and h m can be simultaneously obtained, thus is convenient to use; (2) it is time-saving, just one sorption process in airtight chamber is required; (3) the determination of h m is based on the formaldehyde and VOC concentration data in the test chamber rather than the generally used empirical correlations obtained from the heat and mass transfer analogy, thus is more accurate and can be regarded as a significant improvement. The present method is applied to measure the three parameters by treating the experimental data in the literature, and good results are obtained, which validates the effectiveness of the method. Our new method also provides a potential pathway for measuring h m of semi-volatile organic compounds (SVOC) by using that of VOC.  相似文献   

5.
A simple method is described for the estimation of the Michaelis parameters, Km and Vm, from a single progress curve at a single substrate concentration without the need to follow the reaction to completion. By measuring the substrate concentration and the time when the second derivative is at a minimum, Km and Vm can be easily obtained.  相似文献   

6.
Due to their high porosity and biocompatibility, polyurethane foam (PUF) and cellulose foam were adopted for insect cell immobilization and baculovirus expression. Spodoptera frugiperda (SF-21) cells were grown within the macroporous matrix and then infected by Autographa californica nuclear polyhedrosis virus (AcNPV) which was encoded with human interleukin-5 (hIL-5) gene. An appropriate initial cell loading density and medium circulation velocity determined from the previous study were applied in this actual cell cultivation experiments to obtain a uniform initial and final axial cell distribution. The growth of insect cells and the expression of baculovirus were successful in the macroporous packed bed systems used. The final average cell density in cellulose foam achieved was 5.2×107?cells/cm3 and 4.3×107?cells/cm3 in PUF. Under the conditions of sufficient nutrition and oxygen supplement, the average productivity of hIL-5 in cellulose foam packed bed bioreactor reached 7.2×107 unit/l-day. With 50% fresh medium replacement after viral infection, the average productivity of hIL-5 in PUF packed bed reached 8.4×107 unit/l-day, about two fold than that without any fresh medium replacement at infection.  相似文献   

7.
A methyltransferase, which catalyzes the methylation of luteolin (Km, 16 μM) using S-adenosyl-l-methionine as the methyl donor, has been purified about 38-fold from cell suspension cultures of soybean (Glycine max L., var. Mandarin). The following 3,4-dihydroxy phenolic compounds were also methylated: luteolin 7-O-glucoside (Km, 28 μm), quercetin (Km, 35 μm), eriodictyol (Km, 75 μm), 5-hydroxyferulic acid (Km, 227 μm), dihydroquercetin (Km, 435 μm), and caffeic acid (Km, 770 μm). Rutin and quercetin 3-O-glucoside were poor substrates. Methylation proceeded only in the meta position. The enzyme was unable to catalyze the methylation of p-coumaric acid, m-coumaric acid, ferulic acid, isoferulic acid, sinapic acid, apigenin, or naringenin. While the isoflavones biochanin A and daidzein did not serve as substrates, texasin (6,7-dihydroxy-3′-methoxyisoflavone) was methylated (Km, 35 μm). The methylation of caffeic acid and quercetin showed a pH optimum of 8.6–8.9. The enzyme required Mg2+ ions for maximum activity (approximately 1 mm) and could be totally inhibited by EDTA (10 mm). The Km for S-adenosyl-l-methionine was 11 μm. S-Adenosyl-l-homocysteine inhibited the methylation of luteolin by S-adenosyl-l-methionine.  相似文献   

8.
An analytical solution is obtained for the three-dimensional spatial distribution of potential inside a flat cell, such as the layer of horizontal cells, as a function of its geometry and resistivity characteristics. It was found that, within a very large range of parameter values, the potential is given by [Formula: see text] where r = ρ/ρ0, = z0, ρ = (Ri/Rm)·ρ0, δ = h0; K is a constant; J is the assumed synaptic current; ρ, z are cylindrical coordinates; ρ0 is the radius of the synaptic area of excitation; h is the cell thickness; and Ri, Rm are the intracellular and membrane resistivities, respectively. Formula A closely fits data for the spatial decay of potential which were obtained from the catfish internal and external horizontal cells. It predicts a decay which is exponential down to about 40% of the maximum potential but is much slower than exponential below that level, a characteristic also exhibited by the data. Such a feature in the decay mode allows signal integration over the large retinal areas which have been observed experimentally both at the horizontal and ganglion cell stages. The behavior of the potential distribution as a function of the flat cell parameters is investigated, and it is found that for the range of the horizontal cell thicknesses (10-50 μ) the decay rate depends solely on the ratio Rm/Ri. Data obtained from both types of horizontal cells by varying the diameter of the stimulating spot and for three widely different intensity levels were closely fitted by equation A. In the case of the external horizontal cell, the fit for different intensities was obtained by varying the ratio Rm/Ri; in the case of the internal horizontal cell it was found necessary, in order to fit the data for different intensities, to vary the assumed synaptic current J.  相似文献   

9.
1. Cerebral-cortex mitochondria, after purification by using high-density sucrose solutions, were extracted with Triton X-100. The total hexokinase activity of the intact mitochondria was increased by 50–80% in the Triton extracts. 2. Triton X-100 was removed from mitochondrial extracts by a combination of ammonium sulphate fractionation and DEAE-cellulose chromatography. Mitochondrial hexokinase remained soluble after removal of extractant. 3. The behaviour of solubilized mitochondrial hexokinase was compared with soluble cytoplasmic hexokinase from the same samples of cerebral cortex on identical columns of DEAE-cellulose. Two peaks were eluted from each source of hexokinase. The distribution between hexokinase peaks was similar for the two sources. Peak I (approx. 80% of the total hexokinase) from each was eluted at identical concentrations of potassium chloride and slight differences were observed in the elution profiles for peak II. 4. The purified mitochondrial hexokinase showed the following kinetic properties: peak I, Km(ATP) 0.60mm, Km(glucose) 0.042mm; peak II, Km(ATP) 0.66mm, Km(glucose) 0.043mm. The purified cytoplasmic hexokinase Michaelis constants were: peak I, Km(ATP) 0.56mm, Km(glucose) 0.048mm; peak II, Km(ATP) 0.68mm, Km(glucose) 0.062mm. 5. Although no significant differences between mitochondrial and cytoplasmic hexokinases were noted in chromatographic behaviour or in the kinetic properties studied, the purified mitochondrial enzyme was activated slightly (approx. 20%) by Triton X-100, in contrast with the cytoplasmic enzyme, which was not affected. 6. The results, taken to indicate basic similarity between mitochondrial and cytoplasmic hexokinases, are discussed in relation to the role of the two sources of enzyme in the metabolism of the tissue.  相似文献   

10.
11.
The activation of rabbit intestinal brush-border sucrase in the pH range 4.8 to 9.2 was studied as a function of sucrose concentration and temperature in a metal-free, n-butylamine universal buffer, both in the absence and in the presence of sodium. When sodium was absent, enzyme activation involved the simultaneous loss of two key protons (pK1 of about 5.6), thus yielding a high-affinity, catalytically active enzyme conformation. Inactivation followed when a third key proton (pK2 of about 8.4) was lost. When sodium was present, kinetic analysis in the pH range 4.8 to 7.2 revealed that sodium activation involves distinct effects on the two kinetic parameters, Vm and Km. The Vm parameter seemed to conform to the classical rules of pH-dependent enzyme activation and implicated the release of a single proton whose apparent pK (pK1y, about 5.6) was little affected by sodium. On the contrary, the Km parameter was strongly influenced by sodium. Here, activation of rabbit sucrase seemed to involve release of a different proton whose apparent pK (pK1x also of about 5.6 in the absence of sodium) was strongly shifted to more acid values by saturating sodium concentrations. The functional distinction between the above two protons explains the existence of strong affinity-type activating effects of sodium on rabbit sucrase, previously shown to be pH independent (F. Alvarado and A. Mahmood, 1979, J. Biol. Chem.254, 9534–9541).  相似文献   

12.
V-ATPases are rotary molecular motors that generally function as proton pumps. We recently solved the crystal structures of the V1 moiety of Enterococcus hirae V-ATPase (EhV1) and proposed a model for its rotation mechanism. Here, we characterized the rotary dynamics of EhV1 using single-molecule analysis employing a load-free probe. EhV1 rotated in a counterclockwise direction, exhibiting two distinct rotational states, namely clear and unclear, suggesting unstable interactions between the rotor and stator. The clear state was analyzed in detail to obtain kinetic parameters. The rotation rates obeyed Michaelis-Menten kinetics with a maximal rotation rate (Vmax) of 107 revolutions/s and a Michaelis constant (Km) of 154 μm at 26 °C. At all ATP concentrations tested, EhV1 showed only three pauses separated by 120°/turn, and no substeps were resolved, as was the case with Thermus thermophilus V1-ATPase (TtV1). At 10 μm ATP (⪡Km), the distribution of the durations of the ATP-waiting pause fit well with a single-exponential decay function. The second-order binding rate constant for ATP was 2.3 × 106 m−1 s−1. At 40 mm ATP (⪢Km), the distribution of the durations of the catalytic pause was reproduced by a consecutive reaction with two time constants of 2.6 and 0.5 ms. These kinetic parameters were similar to those of TtV1. Our results identify the common properties of rotary catalysis of V1-ATPases that are distinct from those of F1-ATPases and will further our understanding of the general mechanisms of rotary molecular motors.  相似文献   

13.
We describe here the synthesis and the binding interaction with σ1 and σ2 receptors of a series of new arylcarboxamide derivatives variously substituted on the aromatic portions. Maintaining a partial scaffold of a series of compounds previously synthesized by us, we evaluate the effect of the substitution on σ binding. The synthesized compounds have been tested to estimate their affinity and selectivity toward σ1 and σ2 receptors. Two out of 16 derivatives showed an interesting σ1 affinity (21.2 and 13.6 nM—compounds 2m and 2p) and a good selectivity (Ki2)/Ki1) >140 and >40, respectively).  相似文献   

14.
Nitrate reductase activity from filamentous, heterocyst-forming cyanobacteria showed a biphasic kinetic behavior with respect to nitrate as the variable substrate. Two kinetic components were detected, the first showing a higher affinity for nitrate (Km, 0.05-0.25 mm) and a lower catalytic activity and the second showing a lower affinity for nitrate (Km, 5-25 mm) and a higher (3- to 5-fold) catalytic activity. In contrast, among unicellular cyanobacteria, most representatives studied exhibited a monophasic, Michaelis-Menten kinetic pattern for nitrate reductase activity. Biphasic kinetics remained unchanged with the use of different assay conditions (i.e. cell disruption or permeabilization, two different electron donors) or throughout partial purification of the enzyme.  相似文献   

15.
In pathogenic trypanosomes, trypanothione synthetase (TryS) catalyzes the synthesis of both glutathionylspermidine (Gsp) and trypanothione (bis(glutathionyl)spermidine (T(SH)2)). Here we present a thorough kinetic analysis of Trypanosoma brucei TryS in a newly developed phosphate buffer system at pH 7.0 and 37 °C, mimicking the physiological environment of the enzyme in the cytosol of bloodstream parasites. Under these conditions, TryS displays Km values for GSH, ATP, spermidine, and Gsp of 34, 18, 687, and 32 μm, respectively, as well as Ki values for GSH and T(SH)2 of 1 mm and 360 μm, respectively. As Gsp hydrolysis has a Km value of 5.6 mm, the in vivo amidase activity is probably negligible. To obtain deeper insight in the molecular mechanism of TryS, we have formulated alternative kinetic models, with elementary reaction steps represented by linear kinetic equations. The model parameters were fitted to the extensive matrix of steady-state data obtained for different substrate/product combinations under the in vivo-like conditions. The best model describes the full kinetic profile and is able to predict time course data that were not used for fitting. This system''s biology approach to enzyme kinetics led us to conclude that (i) TryS follows a ter-reactant mechanism, (ii) the intermediate Gsp dissociates from the enzyme between the two catalytic steps, and (iii) T(SH)2 inhibits the enzyme by remaining bound at its product site and, as does the inhibitory GSH, by binding to the activated enzyme complex. The newly detected concerted substrate and product inhibition suggests that TryS activity is tightly regulated.  相似文献   

16.
Membrane potential, Vm, and K+(86Rb+) fluxes have been measured simultaneously on individual cells of Acetabularia mediterranea. During resting state (resting potential approx. ?170 mV) the K+ influx amounts to 0.24–0.6 pmol · cm?2 · s?1 and the K+ efflux to 0.2–1.5 pmol · cm?2 s?1. According to the K+ concentrations inside and outside the cell (40 : 1) the voltage dependent K+ flux (zero at Vm = EK = ?90 mV) is stimulated approx. 40-fold for Vm more positive than EK.It is calculated that during one action potential (temporary depolarization to Vm more positive than EK) a cell looses the same amount of K+, which leaks in during 10–20 min in the resting state (Vm = ?170 mV). Since action potentials occur spontaneously in Acetabularia, they are therefore suggested to have a significant function for the K+ balance of this alga.  相似文献   

17.
The properties of the α1 Na+-K+ pump were compared in Dahl salt-sensitive (DS) and salt-resistant (DR) strains by measuring ouabain-sensitive luxes (mmol/liter cell x hr = FU, Mean ± se) in red blood cells (RBCs) and varying internal ( i ) and external ( o ) Na+ and K+ concentrations. Kinetic parameters of several modes of operation, i.e., Na+/ K+, K+/K+, Na+/Na+ exchanges, were characterized and analyzed for curve-fitting using the Enzfitter computer program. In unidirectional flux studies (n=12 rats of each strain) into fresh cells incubated in 140 mm Na+ + 5 mm K+, ouabain-sensitive K+ influx was substantially lower in the DS than in DR RBCs, while ouabain-sensitive Na+ efflux and Na i were similar in both strains. Thus, the coupling ratio between unidirectional Na+∶K+ fluxes was significantly higher in DS than in DR cells at similar RBC Na+ content. In the presence of 140 mm Na o , activation of ouabain-sensitive K+ influx by K o had a lower K m and V max in DS as estimated by the Garay equation (N=2.70 ± 0.33, K m 0.74 ± 0.09 mm; V max 2.87 ± 0.09 FU) than in DR rats (N=1.23 ± 0.36, K m 2.31 ± 0.16 mm; v max 5.70 ± 0.52 FU). However, the two kinetic parameters were similar following Na o removal. The activation of ouabain-sensitive K+ influx by Na i had significantly lower V max in DS (9.3 ± 0.4 FU) than in DR (14.5 ± 0.6 FU) RBCs but similar K m. These data suggest that the low K+ influx in DS cells is caused by a defect in modulation by Na o and Na i . Na+ efflux showed no differences in Na i activation or trans effects by Na o and K o , thus accounting for the different Na+∶K+ coupling ratio in the Dahl strains. Further evidence for the differences in the coupling of ouabain-sensitive fluxes was found in studies of net Na+ and K+ fluxes, where the net ouabain-sensitive Na+ losses showed similar magnitudes in the two Dahl strains while the net ouabainsensitive K+ gains were significantly greater in the DR than the DS RBCs. Ouabain-sensitive Na+ influx and K+ efflux were also measured in these rat RBCs. The inhibition of ouabain-sensitive Na+ influx by K o was fully competitive for the DS but not for the DR pumps. Thus, for DR pumps, K o could activate higher K+ influx in DR pumps without a complete inhibition of ouabain-sensitive Na+ influx. This behavior is consistent with K o interaction with distinct Na+ and K+ transport sites. In addition, the inhibition of K+ efflux by Na, was different between Dahl strains. Ouabain-sensitive K+ efflux at Na i level of 4.6 mmol/liter cell, was significantly higher in DS (3.86 ± 0.67 FU) than in DR (0.86 ± 0.14 FU) due to a threefold higher K50 for Na i -inhibition 9.66 ± 0.41 vs. 3.09 ± 0.11 mmol/liter cell. This finding indicates that Na+ modulation of K+ transport is altered at both sides of the membrane. The dissociation of Na+ modulatory sites of K+ transport from Na+ transport sites observed in RBCs of Dahl strains suggests that K+ transport by the Na+-K+ pump is controlled by Na+ allosteric sites different from the Na+ transport sites. The alterations in K+ transport may be related to the amino acid substitution (Leu/Gln276) reported for the cDNA of the α1 subunit of the Na+-K+ pump in the DS strain or to post-translational modifications during RBC maturation. These studies were supported by the following grants: NIH (HL-35664, HL-42120, HL-18318, HL-39267, HL-01967). J.R.R. is a Ford Foundation Predoctoral Fellow. A preliminary report of this work was presented at the International Conference on the Na+-K+ pump and 44th Annual Meeting of the Society of General Physiologists held at Woods Hole, MA, September 5–9, 1990, and published as an abstract in the J. Gen. Physiol. 96:70a, 1990.  相似文献   

18.
To assess whether foliar application of K+S as potassium sulfate (K2SO4) could alleviate the adverse effects of salt on sunflower (Helianthus annuus L. cv. SF-187) plants, a greenhouse experiment was conducted. There were two NaCl levels (0 and 150 mM) applied to the growth medium and six levels of K+S as K2SO4 (NS (no spray), WS (spray of water+0.1% Tween 20 solution), 0.5% K+0.21% S, 1.0% K+0.41% S, 1.5% K+0.62% S, and 2.0% K+0.82% S in 0.1% Tween-20 solution) applied two times foliarly to non-stressed and salt-stressed sunflower plants. Salt stress markedly repressed the growth, yield, photosynthetic pigments, water relations and photosynthetic attributes, quantum yield (Fv/Fm), leaf and root K+, Mg2+, P, Ca2+, N as well as K+/Na+ ratios, while it enhanced the cell membrane permeability, and leaf and root Na+ and Cl concentrations. Foliar application of potassium sulfate significantly improved growth, achene yield, photosynthetic and transpiration rates, stomatal conductance, water use efficiency, leaf turgor and enhanced shoot and leaf K+ of the salt-stressed sunflower plants, but it did not improve leaf and root Na+, Cl, Mg2+, P, Ca2+, N as well as K+/Na+ ratios. The most effective dose of K+S for improving growth and achene yield was found to be 1.5% K+0.62% S and 1% K+0.41% S, respectively. Improvement in growth of sunflower plants due to exogenously applied K2SO4 was found to be linked to enhanced photosynthetic capacity, water use efficiency, leaf turgor and relative water content.  相似文献   

19.
Suspensions of 2 to 5% rat thymocytes were incubated at 35 °C in buffered balanced salt solution (pH 7.3) with lactate and β-hydroxybutyrate as fuels. The dependence of 3-O-[Me-3H]methylglucose influx on external and internal 3-O-methylglucose concentrations was studied. Entry was almost rectilinear during the first minute. From the dependence of methylglucose entry (into sugar-free cells) on external methylglucose concentration, we judged the entry Km to be about 7.7 mm and the entry V to be about 0.64 μmol · min?1 · (ml of packed cell volume)?1. Methylglucose inside the cell enhanced influx, hence equilibrium exchange was faster than entry. The dependence of equilibrium exchange on methylglucose concentration (inside and outside being equal) indicated a Km of about 25 mm and a V of about 2.1 μmol · (min)?1 · (ml of cell volume)?1. This effect of internal sugar indicated that entry into sugar-free cells is limited mainly by the return of empty carrier to the outside surface and that loading the carrier on the inside enhances its outward mobility. The Km and V for influx into cells containing 21 mm methylglucose were 5.9 mm and 1.17 μmol · min?1 · (ml of packed cells)?1. The effect of 21 mm internal sugar on lowering the influx Km from about 7.7 mm to about 6 mm was reproducible and contributed to the evaluation of the constants of the transport rate law. It indicated that loading of the carrier at the external surface reduces its mobility, in contrast to the effect of loading on the inside. Mechanical explanations for this behavior are discussed.  相似文献   

20.
Phytase isolated from mung bean cotyledons was purified about 80-fold with a recovery of 28%. The enzyme is stable at 0°, has a pH optimum at 7·5 and optimal temperature of 57°. The energy of activation is approximately 8500 cal/mole between 37° and 57°. Inhibition by Pi has been found to be competitive, the Ki value being 0·40–0·43 × 10−3 M; the Km value with phytate is 0·65 × 10−3 M. Divalent cations are not required for activity. Lower members of inositol phosphates are better substrates, as shown by their Vmax and Km values. When subjected to polyacrylamide gel electrophoresis two bands have been resolved; one (major) corresponds to phytase and the other (minor) to phosphatase and pyrophosphatase activity. Filtration through Biogel P-200 partially resolves phytase from phosphatase and pyrophosphatase. The molecular weight of phytase is approximately 160,000.  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