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 15 毫秒
1.
Three ATP-dependent reactions catalyzed by the inner membrane of rat liver mitochondria and the ATPase reaction catalyzed by purified mitochondrial ATPase (F1), were studied with respect to kinetic properties, substrates specificity, and sensitivity to bicarbonate. The ATP-dependent transhydrogenase reaction (reduction of NADP+ by NADH) catalyzed by inner membrane vesicles displays typical Michaelis-Menten kinetics in both Tris-Cl and Tris-bicarbonate buffers, with Km (ATP) values of 0.035 mM and 0.054 mM respectively. The Vmax of transhydrogenase activity (25 nmol min-1 mg-1) is the same in Tris-bicarbonate or Tris-Cl buffer. ITP and GTP readily substitute for ATP in the transhydrogenase reaction. The ATP-P1 exchange reaction catalyzed by inner membrane vesicles displays typical Michaelis-Menten kinetics in both Tris-Cl and Tris-bicarbonate buffers with Km (ATP) values of 1.0 mM and 1.4 mM respectively. The Vmax of exchange (200 nmol min-1 mg-1) is the same in either buffer. ITP and GTP do not effectively replace ATP in the exchange reaction.  相似文献   

2.
A single mutation, involving the replacement of an arginine residue with histidine to reconstruct a zinc-binding site, suffices to change a catalytically inactive murine carbonic anhydrase-related protein (CARP) to an active carbonic anhydrase with a CO2-hydration turnover number of 1.2 x 104 s-1. Further mutations, leading to a more 'carbonic anhydrase-like' active-site cavity, results in increased activity. A quintuple mutant having His94, Gln92, Val121, Val143, and Thr200 (human carbonic anhydrase I numbering system) shows kcat = 4 x 104 s-1 and kcat/Km = 2 x 107 M-1.s-1, greatly exceeding the corresponding values for carbonic anhydrase isozyme III and approaching those characterizing carbonic anhydrase I. In addition, a buffer change from 50 mM Taps/NaOH to 50 mM 1, 2-dimethylimidazole/H2SO4 at pH 9 results in a 14-fold increase in kcat for this quintuple mutant. The CO2-hydrating activity of a double mutant with His94 and Gln92 shows complex pH-dependence, but the other mutants investigated behave as if the activity (kcat/Km) is controlled by the basic form of a single group with pKa near 7.7. In a similar way to human carbonic anhydrase II, the buffer behaves formally as a second substrate in a ping-pong pattern, suggesting that proton transfer between a zinc-bound water molecule and buffer limits the maximal rate of catalysis in both systems at low buffer concentrations. However, the results of isotope-exchange kinetic studies suggest that proton shuttling via His64 is insignificant in the CARP mutant in contrast with carbonic anhydrase II. The replacement of Ile residues with Val in positions 121 or 143 results in measurable 4-nitrophenyl acetate hydrolase activity. The pH-rate profile for this activity has a similar shape to those of carbonic anhydrase I and II. CD spectra of the double mutant with His94 and Gln92 are variable, indicating an equilibrium between a compact form of the protein and a 'molten globule'-like form. The introduction of Thr200 seems to stabilize the protein.  相似文献   

3.
The rate of ferritin formation in the buffers 4-morpholinepropanesulphonic acid (Mops), 4-morpholineethanesulphonic acid (Mes) and imidazole at pH values from 5.0 to 6.5 is quite similar. However, the rate of iron deposition is much greater in Mops and Mes at pH values above 6.5 than in imidazole. Increasing the concentration of imidazole inhibits ferritin formation and also leads to a transformation in the shape of the kinetic curves observed. This inhibiton is also observed at constant ionic strength but is not found for non-complexing buffers such as Mops. An inhibition of ferritin formation in imidazole and in Mops buffers is also observed with increasing ionic strength. We conclude that the unprotonated form of imidazole inhibits iron deposition, possibly by binding to the active site of the apoferritin molecule. The temperature dependence of iron deposition was examined. An optimum temperature of 50 degrees C was found but the Arrhenius plots were non-linear. On the basis of these and previous results, a kinetic model is developed which accounts well for ferritin formation at pH values below 6.5 and above 7.0 in non-complexing buffers. The model does not account for the kinetics observed at pH values close to neutrality.  相似文献   

4.
We have investigated the steady state and equilibrium kinetic properties of carbonic anhydrase from Neisseria gonorrhoeae (NGCA). Qualitatively, the enzyme shows the same kinetic behaviour as the well studied human carbonic anhydrase II (HCA II). This is reflected in the similar pH dependencies of the kinetic parameters for CO(2) hydration and the similar behaviour of the kinetics of (18)O exchange between CO(2) and water at chemical equilibrium. The pH profile of the turnover number, k(cat), can be described as a titration curve with an exceptionally high maximal value of 1.7 x 10(6) s(-1) at alkaline pH and a pK(a) of 7.2. At pH 9, k(cat) is buffer dependent in a saturable manner, suggesting a ping-pong mechanism with buffer as the second substrate. The ratio k(cat)/K(m) is dependent on two ionizations with pK(a) values of 6.4 and 8.2. However, an (18)O-exchange assay identified only one ionizable group in the pH profile of k(cat)/K(m) with an apparent pK(a) of 6.5. The results of a kinetic analysis of a His66-->Ala variant of the bacterial enzyme suggest that His66 in NGCA has the same function as a proton shuttle as His64 in HCA II. The kinetic defect in the mutant can partially be overcome by certain buffers, such as imidazole and 1,2-dimethylimidazole. The bacterial enzyme shows similar K(i) values for the inhibitors NCO(-), SCN(-) and N(3)(-) as HCA II, while CN(-) and the sulfonamide ethoxzolamide are considerably weaker inhibitors of the bacterial enzyme than of HCA II. The absorption spectra of the adducts of Co(II)-substituted NGCA with acetazolamide, NCO(-), SCN(-), CN(-) and N(3)(-) resemble the corresponding spectra obtained with human Co(II)-isozymes I and II. Measurements of guanidine hydrochloride (GdnHCl)-induced denaturation reveal a sensitivity of the CO(2) hydration activity to the reducing agent tris(2-carboxyethyl)phosphine (TCEP). However, the A(292)/A(260) ratio was not affected by the presence of TCEP, and a structural transition at 2.8--2.9 M GdnHCl was observed.  相似文献   

5.
6.
Y Pocker  N Janji? 《Biochemistry》1988,27(11):4114-4120
In our earlier paper we showed that the rates of CO2 hydration and HCO3- dehydration catalyzed by the high-activity form of mammalian erythrocyte carbonic anhydrase (CA II) were dependent on solution viscosity increase and that the effect was linked to some kind of proton-transfer-related event [Pocker, Y., & Janji?, N. (1987) Biochemistry 26, 2597-2606]. In order to further elucidate the source of the observed viscosity effect, the dependence of kcat and Km for CA II catalyzed HCO3- dehydration at pH 5.90 on sucrose-induced viscosity increase was investigated at several concentrations of 2-(N-morpholino)ethanesulfonic acid (MES) buffer, including the very low buffer concentration region (less than 10 mM) where the proton transfer between the shuttle group on the enzyme and buffer becomes rate limiting. In all examined cases, kcat steadily decreased with added sucrose while Km remained independent of the viscosity increase. The extent to which this reaction was dependent on viscosity was found to be constant, within experimental error, over the entire range of MES buffer concentrations studied (1-20 mM). Furthermore, the viscosity effect was qualitatively and quantitatively the same when an exceptionally large buffer (i.e., bovine serum albumin) was used instead of the more commonly used biological buffer (i.e., MES).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

7.
Purification and characterization of mitochondrial malate dehydrogenase [EC 1.1.1.37] from unfertilized eggs of the sea urchin, Anthocidaris crassispina, are described. The purification method consisted of dextran sulfate fractionation, Blue Dextran Sepharose chromatography, Phenyl-Sepharose hydrophobic chromatography and DEAE-cellulose chromatography. The enzyme was purified 771-fold with a 7% yield from the crude extract. The purified enzyme appeared homogeneous on polyacrylamide gel electrophoresis under both native and denatured conditions. After incubation at 45 degrees C for 50 min, the enzyme lost about 90% of its activity. In the presence of NADH, however, the enzyme was protected against the heat denaturation. The native enzyme had a molecular weight of about 65,000 and probably consisted of two identical subunits. In the reduction of oxaloacetate with NADH, a broad optimum pH ranging from 8.2 to 9.4 was found with 50 mM Tris-HCl and glycine-NaOH buffers. Sodium phosphate buffer apparently activated the enzyme. The apparent Km values for oxaloacetate and NADH were 19 microM and 30 microM, respectively. The optimum pH for malate oxidation with NAD+ was 10.2 in 50 mM NaHCO3-Na2CO3 buffer. The apparent Km values for malate and NAD+ were 7.0 mM and 0.6 mM, respectively. Zinc ion, sulfite ion, p-chloromercuriphenylsulfonate and adenine nucleotides strongly inhibited the enzyme.  相似文献   

8.
Growth of Thiobacillus ferrooxidans in batch culture on 10 mM potassium tetrathionate was optimal at pH 2.5 (specific growth rate, 0.092 h-1). Oxygen electrode studies on resting cell suspensions showed that the apparent Km for tetrathionate oxidation (0.13 to 8.33 mM) was pH dependent, suggesting higher substrate affinity at higher pH. Conversely, oxidation rates were greatest at low pH. High substrate concentrations (7.7 to 77 mM) did not affect maximum oxidation rates at pH 3.0, but produced substrate inhibition at other pH values. Tetrathionate-grown cell suspensions also oxidized thiosulfate at pH 2.0 to 4.0. Apparent Km values (1.2 to 25 mM) were of the same order as for tetrathionate, but kinetics were complex. Continuous culture on growth-limiting tetrathionate at pH 2.5, followed by continuous culture on growth-limiting thiosulfate at pH 2.5, indicated true growth yield values (grams [dry weight] per gram-molecule of substrate) of 12.2 and 7.5, and maintenance coefficient values (millimoles of substrate per gram [dry weight) of organisms per hour) of 1.01 and 0.97 for tetrathionate and thiosulfate, respectively. Yield was increased on both media at low dilution rates by increase in CO2 supply. The apparent maintenance coefficient was lowered without affecting YG, suggesting better energy coupling in CO2-rich environments. Prolonged continuous cultivation on tetrathionate or thiosulfate did not affect the ability of the organism to grow subsequently in ferrous iron medium.  相似文献   

9.
Rates of CO2/HCO-3 exchange, catalyzed by human carbonic anhydrase I (or B) at chemical equilibrium, were estimated from the nuclear magnetic resonance linewidths of 13C-labeled substrates. The results show that the maximal exchange rate constant is independent of pH in the range 5.7-8.0, whereas the apparent substrate dissociation constant depends on pH. Exchange proceeds rapidly in the absence of added buffers, and the addition of buffers has negligible effects on exchange rates. Exchange is equally rapid with 1H2O or 2H2O as solvents. Chloride ions inhibit CO2/HCO-3 exchange competitively. The maximal exchange rates obtained with human carbonic anhydrase I are 50 times slower than those obtained with human isoenzyme II (or C). From a comparison of the exchange kinetics with the steady-state kinetics of CO2 hydration and HCO-3 dehydration it is tentatively concluded that the transfer of H+ between active site and medium proceeds with rates of similar magnitudes in the two isoenzymes, whereas the central catalytic step, the interconversion of enzyme-bound CO2 and HCO-3, is much slower in isoenzyme I than in isoenzyme II.  相似文献   

10.
1. The steady-state kinetics of the interconversion of CO2 and HCO3 catalyzed by human carbonic anhydrase C was studied using 1H2O and 2H2O as solvents. The pH-independent parts of the parameters k(cat) and Km are 3-4 times larger in 1H2O than in 2H2O for both directions of the reaction, while the ratios k(cat)/Km show much smaller isotope effects. With either CO2 or HCO3 as substrate the major pH dependence is observed in k(cat), while Km appears independent of pH. The pKa value characterizing the pH-rate profiles is approximately 0.5 unit larger in 2H2O than in 1H2O. 2. The hydrolysis of p-nitrophenyl acetate catalyzed by human carbonic anhudrase C is approximately 35% faster in 2H2O than in 1H2O. In both solvents the pKa values of the pH-rate profiles are similar to those observed for the CO2-HCO3 interconversion. 3. It is tentatively proposed that the rate-limiting step at saturating concentrations of CO2 or HCO3 is an intramolecular proton transfer between two ionizing groups in the active site. It cannot be decided whether the transformation between enzyme-bound CO2 and HCO3 involves a proton trnasfer or not.  相似文献   

11.
The L-(+)-lactate dehydrogenase (L-lactate:NAD+ oxidoreductase, EC 1.1.1.27) of Streptococcus lactis C10, like that of other streptococci, was activated by fructose 1,6-diphosphate (FDP). The enzyme showed some activity in the absence of FDP, with a pH optimum of 8.2; FDP decreased the Km for both pyruvate and reduced nicotinamide adenine dinucleotide (NADH) and shifted the pH optimum to 6.9. Enzyme activity showed a hyperbolic response to both NADH and pyruvate in all the buffers tried except phosphate buffer, in which the response to increasing NADH was sigmoidal. The FDP concentration required for half-maximal velocity (FDP0.5V) was markedly influenced by the nature of the assay buffer used. Thus the FDP0.5V was 0.002 mM in 90 mM triethanolamine buffer, 0.2 mM in 90 mM tris(hydroxymethyl)aminomethanemaleate buffer, and 4.4 mM in 90 mM phosphate buffer. Phosphate inhibition of FDP binding is not a general property of streptococcal lactate dehydrogenase, since the FDP0.5V value for S. faecalis 8043 lactate dehydrogenase was not increased by phosphate. The S. faecalis and S. lactis lactate dehydrogenases also differed in that Mn2+ enhanced FDP binding in S. faecalis but had no effect on the S. lactis dehydrogenase. The FDP concentration (12 to 15 mM) found in S. lactis cells during logarithmic growth on a high-carbohydrate (3% lactose) medium would be adequate to give almost complete activation of the lactate dehydrogenase even if the high FDP0.5V value found in 90 mM phosphate were similar to the FDP requirement in vivo.  相似文献   

12.
The CO2 hydration activities of cloned human carbonic anhydrase II (carbonate hydro-lyase, EC 4.2.1.1) and variants with Lys, Glu, Gln or Ala replacing His at sequence position 64 have been measured in a variety of different buffers in the pH range 6-9. The variants with Lys-64, Gln-64 and Ala-64 showed non-Michaelis-Menten behavior under some conditions, apparent substrate inhibition being prominent near pH 9. However, asymptotic Michaelis-Menten parameters could be estimated for the limit of low substrate concentrations. All variants show distinct buffer specificities, and imidazole derivatives, Ches and phosphate buffers yield higher kcat values that Bicine, Taps and Mops buffers under otherwise similar conditions. These results are interpreted in terms of different pathways for a rate-limiting proton transfer. In unmodified enzyme, the very high catalytic activity depends on His-64 functioning as an efficient proton transfer group, but this pathway is not available in the variants with Gln-64 and Ala-64. Imidazoles, Ches and phosphate are thought to participate in a metal center-to-buffer proton transfer pathway, whereas Bicine, Taps, Mops and Mes appear to lack this capacity, so that the rate-limiting proton transfer occurs in a metal center-to-bulk water pathway for these variants. The Lys-64 and Glu-64 variants give significantly higher kcat values in Taps, Mops and Mes buffers than the Ala-64 and Gln-64 variants. The pH dependencies of these kcat values are compatible with the hypothesis that Lys-64 and Glu-64 can function as proton transfer groups. Thus, at pH near 9, Lys-64 appears to be only 5-times less efficient than His-64, while Glu-64 is inefficient. At pH 6, Lys-64 is an inefficient proton transfer group, but Glu-64 is only 2-3-times less efficient than His-64. The data indicate that Lys-64 and Glu-64 have pKa values near 8 and below 6, respectively.  相似文献   

13.
The urease proteins of the jack bean (Canavalia ensiformis) and Helicobacter pylori are similar in molecular mass when separated by non-denaturing gradient polyacrylamide gel electrophoresis, both having three main forms. The molecular mass of their major protein form is within the range 440-480 kDa with the other two lesser forms at 230-260 kDa and 660-740 kDa. These forms are all urease active; however, significant kinetic differences exist between the H. pylori and jack bean ureases. Jack bean urease has a single pH optimum at 7.4, whereas H. pylori urease has two pH optima of 4.6 and 8.2 in barbitone and phosphate buffers that were capable of spanning the pH range 3 to 10. The H. pylori Km was 0.6 mM at pH 4.6 and 1.0 mM at pH 8.2 in barbitone buffer, greater than 10.0 mM, and 1.1 mM respectively in phosphate buffer and also greater than 10.0 mM in Tris.HCl at pH 8.2. By comparison, the jack bean urease had a Km of 1.3 mM in Tris.HCl under our experimental conditions. The findings show that the urease activity of H. pylori was inhibited at the pH optimum of 4.6 in the phosphate buffer, but not in the barbitone buffer. This was shown to be due to competitive inhibition by the sodium and potassium ions in the phosphate buffer, not the phosphate ions as suggested earlier. Jack bean urease activity was similarly inhibited by phosphate buffer but again due to the effect of sodium and potassium ions.  相似文献   

14.
A procedure was developed to purify the Streptococcus faecium ATCC 12755 L-alpha-glycerophosphate oxidase. The molecular weight of the purified enzyme was 131,000 and the subunit molecular weight was 72,000. Two moles of FAD were bound/mol of enzyme. Apo-L-alpha-glycerophosphate oxidase displayed physical properties similar to the holoenzyme as judged by electrophoresis in 10% buffer gels at pH 8.5 and by centrifugation in a 5 to 20% linear sucrose gradient. The apoenzyme was completely reactivated by incubation with FAD. L-alpha-Glycerophosphate oxidase was specific for L-alpha-glycerophosphate when compared with several other pohsphorylated glycerol and sugar derivatives. Oxygen was the preferred electron acceptor. At 10 mM DL-alpha-glycerophosphate (below the Km of 26 mM for L-alpha-glycerophosphate), activity was increased from 2.6- to 10-fold by increasing the buffer concentration from 0.01 to 0.1 m. This buffer effect was observed with potassium phosphate and other anionic buffers. In 0.001 m potassium phosphate buffer, pH 7.0, activity was increased by several divalent metal ions, including 10 mM CaCl2 (7.7-fold activation) and 10 mM MgCl, (6.8-fold activation). Fructose 6-phosphate and fructose1-phosphate were inhibitors of the L-alpha-glycerophosphate oxidase.  相似文献   

15.
Interactions of both purified tubulin and microtubule protein (tubulin plus associated proteins) with two commonly used sulfonate buffers were examined. 1,4-Piperazineethanesulfonate (Pipes) and 4-morpholineethanesulfonate (Mes) at high concentrations induce the polymerization of purified tubulin in reactions requiring only buffer, tubulin and GTP. While both reactions were temperature-dependent, cold-reversible and inhibited by GDP, colchicine or Ca2+, there were significant differences between them. Substantially lower tubulin and buffer concentrations were required for Pipes-induced polymerization; and turbidity was much more intense in the Pipes-induced than in the Mes-induced reaction at the same protein concentration. Electron microscopy demonstrated that for the most part typical smooth-walled microtubules were formed in Mes, while aberrant forms were the predominant structures formed in Pipes. When the polymerization of microtubule protein was examined as a function of buffer concentration, biphasic patterns were observed with both Pipes and Mes: polymerization occurred at both low and high, but not intermediate, buffer concentrations. The turbidity observed at high concentrations of Pipes greatly exceeded that at low concentrations. With Mes, equivalent turbidity developed at both high and low buffer concentrations. Although associated proteins copolymerized with tubulin at low buffer concentrations, they were excluded from the polymerized material at high buffer concentrations. Pipes and Mes were compared to sodium phosphate, Tris/HCl and imidazole/HCl buffers at 0.1 M in several polymerization systems using both purified tubulin and microtubule protein. The sulfonate buffers were invariably associated with more vigorous reactions than the other buffers.  相似文献   

16.
The hydrolysis of acetylamino acids by highly purified hog kidney aminoacylase I (N-acylamino acid amidohydrolase, EC 3.5.1.14) was investigated using flow injection analysis to determine reaction rates. We show that the distinctly bell-shaped pH versus activity profiles observed in previous studies do not reflect protonic equilibria in the enzyme, but were created by buffer effects. At low pH, anions such as phosphate, nitrate or chloride markedly increase Km. These effects are reversed at higher pH. In zwitterionic 'Good' buffers (Mes, Mops, and Bicine), maximal velocities are almost independent of pH between 6.5 and 9 for all substrates studied (Ac-LAla, Ac-LGlu, Ac-LMet, Ac-LPhe). Below pH 6.5, the catalytic constants decrease with pH, apparently due to the protonation of a carboxylate with a pKa of 5.5-6. The pH dependence of Km markedly varies among different substates. We conclude that the observed profiles all result from the dissociation of an active-site residue with a pKa of 8-8.5, which we tentatively identify as an active-site cysteine residue. A working model of aminoacylase catalysis is presented that accounts for most of the known facts.  相似文献   

17.
We have developed a protocol to quantify polymer DNA cleavage which replaces the traditional radiolabeling and scintillation counting with fluorescent staining and digital imaging. This procedure offers high sensitivity, speed, and convenience, while avoiding waste and error associated with traditional 32P radiolabeling. This protocol was used to measure cleavage of pBR322 plasmid DNA by EcoRV, a type II restriction enzyme. EcoRV was found to exhibit an order of magnitude difference in binding in two apparently similar buffers used in previous investigations. To determine the origin of this effect, we measured reaction kinetics in buffers of different chemical nature and concentration: Tris, bis-Tris propane, Tes, Hepes, and cacodylate. We found that buffer concentration and identity had significant effects on EcoRV reaction velocity through large changes in specific binding and nonspecific binding (reflected in the Michaelis constant Km and the dissociation constant for nonspecific binding Kns). There were only small changes in Vmax. The source of the buffer effect is the protonated amines common to many pH buffers. These buffer cations likely act as counterions screening DNA phosphates, where both the protonated buffer structure and concentration affect enzyme binding strength. It appears that by choosing anionic buffers or zwitterionic buffers with a buried positive charge, buffer influence on the protein binding to DNA can be largely eliminated.  相似文献   

18.
Cytidine 5'-triphosphate:cytidine 5'-monophosphate-3-deoxy-D-manno-octulosonate cytidylyltransferase (CMP-KDO synthetase) was purified 2,300-fold from frozen Escherichia coli B cells. The enzyme catalyzed the formation of CMP-KDO, a very labile product, from CTP and KDO. No other sugar tested could replace KDO as an alternate substrate. Uridine 5'-triphosphate at pH 9.5 and deoxycytidine 5'-triphosphate at pH 8.0 and 9.5 could be used as alternate substrates in place of CTP. CMP-KDO synthetase required Mg2+ at a concentration of 10.0 mM for optimal activity. The pH optimum was determined to be between 9.6 and 9.3 in tris(hydroxymethyl)aminomethane-acetate or sodium-glycine buffer. This enzyme had an isoelectric point between pH 4.15 and 4.4 and appeared to be a single polypeptide chain with a molecular weight of 36,000 to 40,000. The apparent Km values for CTP and KDO in the presence of 10.0 mM Mg2+ were determined to be 2.0 X 10(-4) and 2.9 X 10(-4) M, respectively, at pH 9.5. Uridine 5'-triphosphate and deoxycytidine 5'-triphosphate had apparent Km values of 8.8 X 10(-4) and 3.4 X 10(-4) M. respectively, at pH 9.5.  相似文献   

19.
The effect of pH on the kinetics of photosynthesis, O(2) inhibition of photosynthesis, and photorespiration was examined with mesophyll cells isolated from soybean (Glycine max [L.] Merr.) leaves. At constant, subsaturating bicarbonate concentration (0.5 mm), O(2) inhibition of photosynthesis increased with increasing pH because high pH shifts the CO(2)-bicarbonate equilibrium toward bicarbonate, thereby reducing the CO(2) concentration. At constant, substrating CO(2) concentrations, cell photorespiration decreased with increasing pH. This was indicated by decreases in the CO(2) compensation concentration, O(2) inhibition of photosynthesis, and glycine synthesis. Km(CO(2)) values for isolated cell photosynthesis and in vitro ribulose-1, 5-diphosphate carboxylase activity decreased with increasing pH, while the Ki(O(2)) for both systems was similar at all pH values. The responses to pH of the corresponding kinetic constants of cell photosynthesis and in vitro RuDP carboxylase with respect to CO(2) and O(2) were identical. This provides additional evidence that the relative rates of photosynthesis and photorespiration in C(3) plants are determined by the kinetic properties of RuDP carboxylase.  相似文献   

20.
T J Wheeler  J D Whelan 《Biochemistry》1988,27(5):1441-1450
It has been claimed that the Km for infinite-cis uptake of glucose in human erythrocytes is so low that the carrier model for transport must be rejected. We redetermined this parameter for three experimental conditions and found instead that the Km values were in good agreement with the model. For each of a variety of cis glucose concentrations, cells were preequilibrated with various concentrations of glucose, and the apparent Km was determined as the intracellular concentration reducing the initial rate of net uptake by half. The dependence of the apparent Km values on the cis glucose was as predicted by the carrier model; the infinite-cis Km was determined from both this concentration dependence and the extrapolated value at infinite cis glucose. The resulting values were 15 mM for fresh blood at 0 degrees C, 39 mM for outdated blood at 0 degrees C, and 11 mM for outdated blood at 25 degrees C. Previous measurements of the Km at room temperature yielded values of 2-3 mM. These earlier studies used a time course procedure that indicated rapid changes in rates during the initial 10 s of uptake but did not directly measure such changes. We examined the uptake of 60 mM glucose at 20 degrees C into cells containing 0 and 5 mM glucose; rapid changes in rates were not observed in the first few seconds, and the time courses were more consistent with our higher Km values. Our new values, together with other initial rate measurements in the literature, support the adequacy of the carrier model to account for the kinetics of glucose transport in human erythrocytes.  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