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 455 毫秒
1.
Yeast phosphofructokinase (PFK) exists in two forms, an ATP-sensitive form, PFKs, and a desensitized form, PFKd(MgF+). PFKs exhibits sigmoidal kinetics with respect to Fru-6-P, whereby the S0.5, Fru-6-P is determined by [ATP]. This form of PFK is inhibited by ATP and citrate and allosterically activated by Fru-6-P and AMP. NH4+ activates PFKs and enhances its affinity for substrate Fru-6-P (1–3).PFKd(MgF+) in contrast is not inhibited by ATP and citrate, nor activated by Fru-6-P and AMP. Kinetics of the reaction with PFKd(MgF+) with respect to Fru6-P are hyperbolic, with Km = 14?15 of S0.5, Fm-6-P for PFKs. NH4+ strongly activates this form.In terms of the model of Monod et al. (4), PFKd(MgF+) corresponds to a fixed R-conformation, while PFKs is a limiting T-conformation.  相似文献   

2.
Luit Slooten  Adriaan Nuyten 《BBA》1984,766(1):88-97
(1) Rates of ATP synthesis and ADP-arsenate synthesis catalyzed by Rhodospirillum rubrum chromatophores were determined with the firefly luciferase method and by a coupled enzyme assay involving hexokinase and glucose-6-phosphate dehydrogenase. (2) Vm for ADP-arsenate synthesis was about 2-times lower than Vm for ATP-synthesis. With saturating [ADP], K(Asi) was about 20% higher than K(Pi). With saturating [anion], K(ADP) was during arsenylation about 20% lower than during phosphorylation. (3) Plots of 1v vs. 1[substrate] were non-linear at low concentrations of the fixed substrate. The non-linearity was such as to suggest a positive cooperativity between sites binding the variable substrate, resulting in an increased VmKm ratio. High concentrations of the fixed substrate cause a similar increase in VmKm, but abolish the cooperativity of the sites binding the variable substrate. (4) Low concentrations of inorganic arsenate (Asi) stimulate ATP synthesis supported by low concentrations of Pi and ADP about 2-fold. (5) At high ADP concentrations, the apparent Ki of Asi for inhibition of ATP-synthesis was 2–3-times higher than the apparent Km of Asi for arsenylation; the apparent Ki of Pi for inhibition of ADP-arsenate synthesis was about 40% lower than the apparent Km of Pi for ATP synthesis. (6) The results are discussed in terms of a model in which Pi and Asi compete for binding to a catalytic as well as an allosteric site. The interaction between these sites is modulated by the ADP concentration. At high ADP concentrations, interaction between these sites occurs only when they are occupied with different species of anion.  相似文献   

3.
The activity of leukocyte glycogen synthetase in a freshly prepared homogenate is almost completely in the b form. Incubation of the homogenate at 30°C caused a time dependent increase in the activity measured in the absence of G-6-P (b to a conversion). The Ka for G-6-P decreased from 0.7 to 0.01 mM. Freezing of the homogenate resulted in a complete loss of the capacity for activation. These results demonstrate that glycogen synthetase from leukocytes of normal human subjects can be converted in vitro to a form, which is almost independent of G-6-P for activity.  相似文献   

4.
The addition of glucagon to hepatocytes in primary culture produced a rapid and sustained increase in the Km (1.27 mM phosphoenol pyruvate) of pyruvate kinase. The low Km (0.4 mM) form of the enzyme was seen when cells were retreated with insulin, demonstrating a short-term regulation mechanism. Injections of insulin, glucagon or glucagon followed by insulin demonstrated that a similar mechanism occurs invivo. Results from longer times after injection indicated that another mechanism occurs when altered activity was the result of changes in Vmax and not Km. Thus, a dual mechanism for regulation of pyruvate kinase occurs. A rapid responding system functions by modification of the enzyme, while a long-term system functions by altering the rate of synthesis, thus changing the amount of enzyme present.  相似文献   

5.
Inorganic pyrophosphate:D-fructose-6-phosphate 1-phosphotransferase from mung beans (Phaseolusaureus Roxb.) was activated markedly by D-fructose 2,6-bisphosphate, with a KA of about 50 nM. The enzyme exhibited hyperbolic kinetics both in the absence and presence of the activator. D-Fructose 2,6-bisphosphate (1 μM) decreased the Km for D-fructose 6-phosphate 67-fold (from 20 mM to 0.3 mM) and increased the Vmax 15-fold; these two effects combined to give a 500-fold activation at 0.3 mM D-fructose 6-phosphate. In contrast, ATP:D-fructose 6-phosphate 1-phosphotransferase from the same source was found not to be affected by D-fructose 2,6-bisphosphate.A natural activator for inorganic pyrophosphate:D-fructose 6-phosphate 1-phosphotransferase was isolated from mung-bean extracts and identified as D-fructose 2,6-bisphosphate.  相似文献   

6.
The effects of treating nitrogen-starved cultures of Escherichia coli W4597 (K) with various doses of 2,4-dinitrophenol include increases in the rates of glucose utilization, decreases in ATP and glucose-6-P and maintenance of the level of fructose-1, 6-P2. A quantitative correlation was observed between the increases in the rates of glucose utilization and decreases in glucose-6-P in agreement with the observation made in vitro that glucose-6-P inhibits glucose transport in E. coli. A quantitative correlation was also observed between glucose-6-P and ATP indicating that the fall in glucose-6-P is effected by the fall in ATP which indirectly signals increased glucose utilization and increased ATP production.  相似文献   

7.
The regulatory properties of citrate on the activity of phosphofructokinase (PFK) purified from rat-kidney cortex has been studied. Citrate produces increases in the K0.5 for Fru-6-P and in the Hill coefficient as well as a decrease in the Vmax of the reaction without affecting the kinetic parameters for ATP as substrate. ATP potentiates synergistically the effects of citrate as an inhibitor of the enzyme. Fru-2,6-P2 and AMP at concentrations equal to Ka were not able to completely prevent citrate inhibition of the enzyme. Physiological concentrations of ATP and citrate produce a strong inhibition of renal PFK suggesting that may participate in the control of glycolysisin vivo.Abbreviations PFK 6-Phosphofructo-1-kinase (EC 2.7.1.11) - Fru-6-P Fructose 6-phosphate - Fru-2,6-P2 Fructose 2,6-bisphosphate  相似文献   

8.
Quercetin inhibited a dog kidney (Na+ + K+)-ATPase preparation without affecting Km for ATP or K0.5 for cation activators, attributable to the slowly-reversible nature of its inhibition. Dimethyl sulfoxide, a selector of E2 enzyme conformations, blocked this inhibition, while the K+-phosphatase activity was at least as sensitive to quercetin as the (Na+ + K+)-ATPase activity, all consistent with quercetin favoring E1 conformations of the enzyme. Oligomycin, a rapidly-reversible inhibitor, decreased the Km for ATP and the K0.5 for cation activators, and its inhibition was also diminished by dimethyl sulfoxide. Although oligomycin did not inhibit the K+-phosphatase activity under standard assay conditions, a reaction presumably catalyzed by E2 conformations, its effects are nevertheless accommodated by a quantitative model for that reaction depicting oligomycin as favoring E1 conformations. The model also accounts quantitatively for effects of both dimethyl sulfoxide and oligomycin on Vmax, Km for substrate, and K0.5 for K+, as well as for stimulation of phosphatase activity by both these reagents at low K+ but high Na+ concentrations.  相似文献   

9.
John D. Mills  Peter Mitchell 《BBA》1984,764(1):93-104
Thiol modulation of the chloroplast protonmotive ATPase (CF0-CF1) by preillumination of broken chloroplasts in the presence of dithiothreitol (or preillumination of intact chloroplasts in the absence of added thiols) had the following effects on photophosphorylation. (1) When assayed at pH 8 and saturating light, the initial rate of photophosphorylation was increased by 10–40%. There was an accompanying increase in the rate of coupled electron transport with no significant change in the overall P2e ratio. (2) On lowering the pH of the assay medium to pH 7, the stimulatory effect of thiol modulation on photophosphorylation and coupled electron flow was enhanced. At pH 7, there was also a small increase in P2e ratio. (3) Addition of a non-saturating amount of uncoupler to the assay medium enhanced the stimulatory effect of thiol modulation on photophosphorylation. In the presence of 1 mM NH4Cl, there was only a small increase in coupled electron flow and a correspondingly larger increase in P2e ratio. (4) Lowering the light intensity, or inhibiting electron transport, diminished the stimulatory effect of thiol modulation on photophosphorylation, coupled electron transport and P2e ratio. (5) Under all the above conditions, the ΔpH maintained across the thylakoid membrane was lower after thiol modulation, even when photophosphorylation markedly increased in rate. (6) Thiol modulation of CF0-CF1 increased the observed Michaelis constant for ADP (Km(ADP)) and the apparent maximum rate (Vapp of photophosphorylation by the same factor, so that ratio VappKm was not altered. VappKm was also unaffected by changing the medium pH, but was significantly decreased upon addition of uncouplers to the medium. These results indicate that the observed rate of ATP synthesis catalysed by thiol demodulated chloroplasts is limited kinetically by the fraction (α) of enzyme molecules that are active during photophosphorylation. A model based on a dual pH optimum requirement for activation of CF0-CF1 is presented to explain the dependence of α on ΔpH. Thiol modulation of CF0-CF1 is proposed to stimulate photophosphorylation by causing the enzyme to become active over a lower range of ΔpH, thereby reducing the kinetic limitation on ATP synthesis imposed by the activation process.  相似文献   

10.
A kynureninase-type enzyme was isolated from adult mouse liver. With kynurenine as the substrate, this enzyme has a Km of 300 μM; when the substrate is hydroxykynurenine, the Km is 6 μM. We conclude that this enzyme is an hydroxykynureninase. No enzyme which we could characterize as a kynureninase was found in this preparation. This suggests that tryptophan metabolism in the mouse occurs primarily through pathways that use hydroxykynurenine rather than kynurenine. Preliminary studies indicate that the enzyme is inhibited by its reaction product, hydroxyanthranilate, which is an intermediate in the synthesis of NAD. Such control of the hydroxykynureninase reaction may be of physiological importance in regulating the synthesis of NAD and/or in preventing the accumulation of hydroxyanthranilate, a putative carcinogen.  相似文献   

11.
Hydrolysis of benzyloxycarbonyl-GlyGlyPhe by nitro(Tyr 248)carboxypeptidase A over the pH range 4.88–8.04 has been examined. The nitroenzyme retains appreciable activity near pH 6.5, and the limiting value of Km is scarcely affected. The peptidase activity has a pH dependence characterized by the following parameters: pKE1 of 6.37 ± 0.19 and pKE2 of 6.60 ± 0.17 in kcatKm, and apparent pK of 5.59 ± 0.06 in Kcat. A spectroscopic pK of 6.75 ± 0.01, attributable to the nitro-Tyr 248 residue, has been determined. This correlates with the base-limb pKE2 in the kcatKm profile, which appears to be shifted from a higher value, pKE2 of 9.0, for the native enzyme. The single (acid-limb) pK which characterizes the kcat profile of the native enzyme is also found to be perturbed to a lesser extent by nitration. A kinetically competent reverse protonation mechanism, based on chemical modification and crystallographic evidence for the enzyme, is described.  相似文献   

12.
The parameters Km and kcat were determined for 16 methyl hippurates (CH3OCOCH2NHCOC6H4-X) hydrolyzed by papain. A simple linear relationship is found between log 1Km and the hydrophobic substituent constant π. It is found that log kcat is parabolically related to π. The results with papain are compared with results obtained by Hawkins and Williams with the enzyme bromelain. The two enzymes behave in a similar fashion.  相似文献   

13.
The effect of pH on the kinetic parameters for the chloroperoxidase-catalyzed N-demethylation of N,N-dimethylaniline supported by ethyl hydroperoxide was investigated from pH 3.0 to 7.0. Chloroperoxidase was found to be stable throughout the pH range studied. Initial rate conditions were determined throughout the pH range. The Vmax for the demethylation reaction exhibited a pH optimum at approximately 4.5. The Km for N,N-dimethylaniline increased with decreasing pH, while the Km for ethyl hydroperoxide varied in a manner paralleling Vmax. Comparison of the VmaxKm values for N,N-dimethylaniline and ethyl hydroperoxide indicated that the interaction of N,N-dimethylaniline with chloroperoxidase compound I was rate-limiting below pH 4.5, while compound I formation was rate-limiting above pH 4.5. The log of the VmaxKm for ethyl hydroperoxide was independent of pH, indicating that chloroperoxidase compound I formation is not affected by ionizations in this pH range. The plot of the log of the VmaxKm for N,N-dimethylaniline versus pH indicated an ionization on compound I with a pK of approximately 6.8. The plot of the log of the Vmax versus pH indicated an ionization on the compound I-N,N-dimethylaniline complex, with a pK of approximately 3.1. The results show that chloroperoxidase can demethylate both the protonated and neutral forms of N,N-dimethylaniline (pK approximately 5.0), suggesting that hydrophobic binding of the arylamine substrate is more important in catalysis than ionic bonding of the amine moiety. For optimal catalysis, a residue in the chloroperoxidase compound I-N,N-dimethylaniline complex with a pK of approximately 3.1 must be deprotonated, while a residue in compound I with a pK of approximately 6.8 must be protonated.  相似文献   

14.
Na+-ATPase activity of a dog kidney (Na+ + K+)-ATPase enzyme preparation was inhibited by a high concentration of NaCl (100 mM) in the presence of 30 μM ATP and 50 μM MgCl2, but stimulated by 100 mM NaCl in the presence of 30 μM ATP and 3 mM MgCl2. The K0.5 for the effect of MgCl2 was near 0.5 mM. Treatment of the enzyme with the organic mercurial thimerosal had little effect on Na+-ATPase activity with 10 mM NaCl but lessened inhibition by 100 mM NaCl in the presence of 50 μM MgCl2. Similar thimerosal treatment reduced (Na+ + K+)-ATPase activity by half but did not appreciably affect the K0.5 for activation by either Na+ or K+, although it reduced inhibition by high Na+ concentrations. These data are interpreted in terms of two classes of extracellularly-available low-affinity sites for Na+: Na+-discharge sites at which Na+-binding can drive E2-P back to E1-P, thereby inhibiting Na+-ATPase activity, and sites activating E2-P hydrolysis and thereby stimulating Na+-ATPase activity, corresponding to the K+-acceptance sites. Since these two classes of sites cannot be identical, the data favor co-existing Na+-discharge and K+-acceptance sites. Mg2+ may stimulate Na+-ATPase activity by favoring E2-P over E1-P, through occupying intracellular sites distinct from the phosphorylation site or Na+-acceptance sites, perhaps at a coexisting low-affinity substrate site. Among other effects, thimerosal treatment appears to stimulate the Na+-ATPase reaction and lessen Na+-inhibition of the (Na+ + K+)-ATPase reaction by increasing the efficacy of Na+ in activating E2-P hydrolysis.  相似文献   

15.
The two aspartate aminotransferase isoenzymes, AAT-P1 and AAT-P2, found in the plant cytosol fraction of lupine nodules have been separated and partially purified. Both isoenzymes showed a broad pH optimum between 7.0 and 9.5 and demonstrated high substrate specificity. Molecular weights determined by gel filtration chromatography were 105,000 and 96,000 for AAT-P1 and AAT-P2, respectively. AAT-P1 demonstrated a subunit molecular weight of 47,000 and AAT-P2, 45,000. Both isoenzymes showed oxaloacetate substrate inhibition and gave similar Km values for aspartate (2.2 and 2.6 mm) and 2-oxoglutarate (0.26 and 0.2 mm) for AAT-P1 and AAT-P2, respectively. However, AAT-P2 showed a fivefold lower Km for oxaloacetate (0.02 mm) and a 1.8-fold lower Km for glutamate (12 mm) than did AAT-P1 for these substrates. The parallel nature of the 1Vvs1[S] plots together with the product inhibition kinetics were consistent with a ping-pong-bi-bi mechanism of action. The results are discussed in terms of the possible physiological significance of these plant aspartate aminotransferases in ammonia assimilation in lupine nodules.  相似文献   

16.
Showdomycin inhibited pig brain (Na+ + K+)-ATPase with pseudo first-order kinetics. The rate of inhibition by showdomycin was examined in the presence of 16 combinations of four ligands, i.e., Na+, K+, Mg2+ and ATP, and was found to depend on the ligands added. Combinations of ligands were divided into five groups in terms of the magnitude of the rate constant; in the order of decreasing rate constants these were: (1)Na+ + Mg2+ + ATP, (2) Mg2+, Mg2+ + K+, K+ and none, (3) Na+ + Mg2+, Na+, K+ + Na+ and Na+ + K+ + Mg2+, (4) Mg2+ + K+ + ATP, K+ + ATP and Mg2+ + ATP, (5)K+ + Na+ + ATP, Na+ + ATP, Na+ + ATP, Na+ + K+ + Mg2+ + ATP and ATP. The highest rate was obtained in the presence of Na+, Mg2+ and ATP. The apparent concentrations of Na+, Mg2+ and ATP for half-maximum stimulation of inhibition (K0.5s) were 3 mM, 0.13 mM and 4μM, respectively. The rate was unchanged upon further increase in Na+ concentration from 140 to 1000 mM. The rates of inhibition could be explained on the basis of the enzyme forms present, including E1, E2, ES, E1-P and E2-P, i.e., E2 has higher reactivity with showdomycin than E1, while E2-P has almost the same reactivity as E1-P. We conclude that the reaction of (Na+ + K+)-ATPase proceeds via at least four kinds of enzyme form (E1, E2, E1 · nucleotide and EP), which all have different conformations.  相似文献   

17.
ω-Hydroxyfatty acid:NADP oxidoreductase, an enzyme involved in suberin biosynthesis, is induced by wounding potato tubers. Initial velocity and product inhibition studies with the purified enzyme suggested an ordered sequential mechanism, where NADPH is added first, followed by 16-oxohexadecanoate, and NADP is released after 16-hydroxyhexadecanoate. Substrate inhibition by NADPH was observed at concentrations higher than 0.2 mm. The inhibitory NADPH molecule competes with 16-oxohexadecanoate, indicating that it forms a dead-end complex with the E-NADPH form of the enzyme. The kinetics for the NADPH inhibition suggested that n > 1 in the rate equation v = V[NADPH](Km + [NADPH]+ [NADPH]n+1Ki); i.e., more than two NADPH molecules bind to enzyme. The Km for 16-oxohexadecanoate did not change from pH 7.5 to 9.0 but increased about 10-fold from pH 9.0 to 10.0, whereas the Km for NADPH and hexadecanal did not vary significantly in this pH range. Phenylglyoxal inactivated the enzyme; NADPH and AMP (which competes with NADPH; Ki = 1.1 mM) provided protection against such inactivation. Diethylpyrocarbonate also caused inactivation which was reversed by hydroxylamine; NADPH but not AMP protected the enzyme from this inhibition. Pyridoxal-5′-phosphate reversibly inactivated the enzyme and NaBH4 reduction of the pyridoxal phosphate-treated enzyme resulted in irreversible inhibition; a combination of NADPH and ω-oxo C16 acid provided protection against such inactivation. As the chain length of alkanals increased from C3 to C8, the Km for the substrate decreased drastically from 7000 to 90μm and a further increase in chain length from C8 to C20 resulted in only a small decrease in Km. The Km and V for 8-oxooctanoate and 10-oxodecanoate are compared with the values obtained for 16-oxohexadecanoate. Based on these results, it is proposed that arginine acts as the binding site for NADPH, a hydrophobic crevice with lysine at the bottom forms the binding site for 16-oxohexadecanoate and histidine participates in the reaction as the proton donor.  相似文献   

18.
An ATPase is demonstrated in plasma membrane fractions of goldfish gills. This enzyme is stimulated by Cl? and HCO3?, inhibited by SCN?.Biochemical characterization shows that HCO3? stimulation (Km = 2.5 mequiv./l) is specifically inhibited in a competitive fashion by SCN? (Ki = 0.25 mequiv./l). The residual Mg2+-dependent activity is weakly is weakly affected by SCN?.In the microsomal fraction chloride stimulation of the enzyme occurs in the presence of HCO3? (Kmfor chloride = 1 mequiv./l); no stimulation is observed in the absence of HCO3?. Thiocyanate exhibits a mixed type of inhibition (Ki = 0.06 mequiv./l) towards the Cl? stimulation of the enzyme.Bicarbonate-dependent ATPase from the mitochondrial fraction is stimulated by Cl?, but this enzyme has a relatively weak affinity for this substrate (Km = 14 mequiv./l).  相似文献   

19.
(1) The t12 for 1.3 mM D-allose uptake and efflux in insulin-stimulated adipocytes is 1.7 ± 0.1 min. In the absence of insulin mediated uptake of D-allose is virtually eliminated and the uptake rate (t12 = 75.8 ± 4.99 min) is near that calculated for nonmediated transport. The kinetic parameters for D-allose zero-trans uptake in insulin-treated cells are Kztoi = 271.3 ± 34.2 mM, Vztoi = 1.15 ± 0.12 mM · s?1. (2) A kinetic analysis of the single-gate transporter (carrier) model interacting with two substrates (or substrate plus inhibitor) is presented. The analysis shows that the heteroexchange rates for two substrates interacting with the transporter are not unique and can be calculated from the kinetic parameters for each sugar acting alone with the transporter. This means that the equations for substrate analogue inhibition of the transport of a low affinity substrate such as D-allose can be simplified. It is shown that for the single gate transporter the Ki for a substrate analogue inhibitor should equal the equilibrium exchange Km for this analogue. (3) Analogues substituted at C-1 show a fused pyranose ring is accepted by the transporter. 1-Deoxy-D-glucose is transported but has low affinity for the transporter. High affinity can be restored by replacing a fluorine in the β-position at C-1. The Ki for d-glucose = 8.62 mM; the Ki for β-fluoro-d-glucose = 6.87 mM. Replacing the ring oxygen also results in a marked reduction in affinity. The Ki for 5-thio-d-glucose = 42.1 mM. (4) A hydroxyl in the gluco configuration at C-2 is not required as 2-deoxy-d-galactose (Ki = 20.75 mM) has a slightly higher affinity than d-galactose (Ki = 24.49 mM). A hydroxyl in the manno configuration at C-2 interferes with transport as d-talose (Ki = 35.4 mM) has a lower affinity than d-galactose. (5) d-Allose (Km = 271.3 mM) and 3-deoxy-d-glucose (Ki = 40.31 mM) have low affinity but high affinity is restored by substituting a fluorine in the gluco configuration at C-3. The Ki for 3-fluoro-d-glucose = 7.97 mM. (6) Analogues modified at C-4 and C-6 do not show large losses in affinity. However, 6-deoxy-d-glucose (Ki = 11.08 mM) has lower affinity than d-glucose and 6-deoxy-d-galactose Ki = 33.97 mM) has lower affinity than d-galactose. Fluorine substitution at C-6 of d-galactose restores high affinity. The Ki for 6-fluoro-d-galactose = 6.67 mM. Removal of the C-5 hydroxymethyl group results in a large affinity loss. The Kid-xylose = 45.5 mM. The Ki for l-arabinose = 49.69 mM. (7) These results indicate that the important hydrogen bonding positions involved in sugar interaction with the insulin-stimulated adipocytes transporter are the ring oxygen, C-1 and C-3. There may be a weaker hydrogen bond to C-6. Sugar hydroxyls in non-gluco configurations may sterically hinder transport.  相似文献   

20.
(1) Treatment of (Na+ + K+)-ATPase from rabbit kidney outer medulla with the γ-35S labeled thio-analogue of ATP in the presence of Na+ + Mg2+ and the absence of K+ leads to thiophosphorylation of the enzyme. The Km value for [γ-S]ATP is 2.2 μM and for Na+ 4.2 mM at 22°C. Thiophosphorylation is a sigmoidal function of the Na+ concentration, yielding a Hill coefficient nH = 2.6. (2) The thio-analogue (Km = 35 μM) can also support overall (Na+ + K+)-ATPase activity, but Vmax at 37°C is only 1.3 γmol · (mg protein)? · h?1 or 0.09% of the specific activity for ATP (Km = 0.43 mM). (3) The thiophosphoenzyme intermediate, like the natural phosphoenzyme, is sensitive to hydroxylamine, indicating that it also is an acylphosphate. However, the thiophosphoenzyme, unlike the phosphoenzyme, is acid labile at temperatures as low as 0°C. The acid-denatured thiophosphoenzyme has optimal stability at pH 5–6. (4) The thiophosphorylation capacity of the enzyme is equal to its phosphorylation capacity, indicating the same number of sites. Phosphorylation by ATP excludes thiophosphorylation, suggesting that the two substrates compete for the same phosphorylation site. (5) The (apparent) rate constants of thiophosphorylation (0.4 s?1 vs. 180 s?1), spontaneous dethiophosphorylation (0.04 s?1 vs. 0.5 s?1) and K+-stimulated dethiophosphorylation (0.54 s?1 vs. 230 s?1) are much lower than those for the corresponding reactions based on ATP. (6) In contrast to the phosphoenzyme, the thiophosphoenzyme is ADP-sensitive (with an apparent rate constant in ADP-induced dethiophosphorylation of 0.35 s?1, KmADP = 48 μM at 0.1 mM ATP) and is relatively K+-insensitve. The Km for K+ in dethiophosphorylation is 0.9 mM and in dephosphorylation 0.09 mM. The thiophosphoenzyme appears to be for 75–90% in the ADP-sensitive E1-conformation.  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