首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
2.
Glycerate was found to effect photosynthetic O2 evolution in wheat chloroplasts by its conversion to triose phosphate and by influencing the rate of photosynthesis through the reductive pentose phosphate pathway. In the absence of bicarbonate, the photosynthetic O2 evolution with glycerate was low (10 to 25 mumol mg chlorophyll-1 h-1), and only about 15% of the rate of bicarbonate-dependent O2 evolution under optimum conditions. This corresponds to a rate of glycerate conversion to triose phosphate of 20 to 50 mumol mg chlorophyll-1 h-1, which appears sufficient to accommodate flux through the glycolate pathway in vivo. Pi was required for this glycerate-dependent O2 evolution; rates remained relatively constant between 0.1 and 40 mM Pi, and proceeded with little lag upon illumination (less than 0.5 min). Evidence for O2 evolution due to glycerate conversion to triose phosphate could be conclusively demonstrated by addition of glycolaldehyde, an inhibitor of the regenerative phase of photosynthesis, which prevents CO2 fixation. The effect of glycerate on photosynthesis in the presence of bicarbonate was determined by measuring both photosynthetic O2 evolution and 14CO2 fixation at varying Pi concentrations. Low concentrations of glycerate (micro- to millimolar levels) prevented inhibition of photosynthesis by Pi. With 1 mM bicarbonate and pH 8.2, which is favorable for glycolate synthesis, maximum rates of photosynthesis were obtained at low Pi (25 microM), whereas strong inhibition of photosynthesis occurred at only 0.2 mM Pi. Addition of glycerate relieved the inhibition of photosynthesis by Pi, indicating the possible importance of glycerate metabolism in the chloroplast under photorespiratory conditions. The initiation of photosynthesis by glycerate at inhibitory Pi levels occurred with little reduction in the ratio of CO2 fixed/O2 evolved, and the main effect of glycerate was on carbon assimilation. While the basis for the beneficial effect of glycerate on CO2 assimilation under moderate to high Pi levels is uncertain, it may increase the concentration of 3-phosphoglycerate (PGA) in the chloroplast, and thus make conditions more favorable for induction of photosynthesis and reduction of PGA to triose phosphate.  相似文献   

3.
The ATPase (EC 3.6.1.3) activity of 30 S dynein from Tetrahymena cilia was remarkably stimulated by porcine brain tubulin at pH 10. The activity increased with increasing concentration of tubulin until the molar ratio of tubulin dimer to 30 S dynein reached approx. 10. The optimum of the ATPase activity of 30 S dynein in the presence of tubulin was 1-2 mM for MgCl2 and 2 mM for CaCl2. Increasing ionic strength gradually inhibited the stimulation effects of tubulin. Activation energies of 30 S dynein in the presence and absence of tubulin were almost the same. At the temperatures beyond 25 degrees C stimulation effects of tubulin disappeared. ATP was a specific substrate even in the presence of tubulin. In kinetic investigations parallel reciprocal plots were observed in a constant ratio of divalent cations to ATP of 2, indicating that tubulin was less tightly bound to 30 S dynein in the presence of ATP than the absence. The similar results were obtained at pH 8.2. 14 S dynein and the 12 S fragment which have poor ability to recombine with outer fibers were also activated with brain tubulin.  相似文献   

4.
An improved method was devised to purify ribulose 1,5-bisphosphate carboxylase/oxygenase (RuBisCO) with high specific activity (2.1 mumol of CO2 fixed/mg protein/min) from Euglena gracilis Z. The purified enzyme stored at -80 degrees C required treatment with dithiothreitol for full activity. The dithiothreitol-treated RuBisCO was activated by 12 mM NaHCO3 and 20 mM MgCl2, and the activated state was stable at least for 60 min in the presence of 4 mM ethylenediaminetetraacetate. The form of inorganic carbon fixed by the Euglena enzyme was CO2, as for the plant enzymes. The carboxylase reaction proceeded linearly with time for at least 8 min. The optimum pH for this reaction was 7.8 to 8.0. The carboxylase activity increased with increasing temperature up to 50 degrees C. The activation energy for the carboxylation reaction was 10.0 kcal/mol. The Michaelis constants of Euglena RuBisCO were 30.9 microM for CO2, 560 microM for O2, and 10.5 microM for ribulose 1,5-bisphosphate. Mathematical comparison between the photosynthesis rate predicted from these enzymatic properties and the observed rate suggested that there is no CO2-concentrating mechanism in E. gracilis.  相似文献   

5.
Experiments performed at micromolar concentrations of inorganic phosphate support the conclusion that liver phosphofructokinase 2 would be completely inactive in the absence of inorganic phosphate or arsenate. The concentration of inorganic phosphate that allowed half-maximal activity decreased with increasing pH, being approximately 0.11 mM at pH 6.5 and 0.05 mM at pH 8. The effect of phosphate was to increase V and to decrease Km for fructose 6-phosphate, without affecting Km for ATP. Citrate and P-enolpyruvate inhibited the enzyme non-competitively with fructose 6-phosphate and independently of the concentration of inorganic phosphate. Phosphorylation of the enzyme by the catalytic subunit of cyclic-AMP-dependent protein kinase did not markedly modify the phosphate requirement and its effect of inactivating phosphofructokinase 2 could not be counteracted by excess phosphate. A nearly complete phosphate dependency was also observed with phosphofructokinase 2 purified from Saccharomyces cerevisiae or from spinach leaves. By contrast, the fructose 2,6-bisphosphatase activity of the liver bifunctional enzyme was not dependent on the presence of inorganic phosphate. Phosphate increased this activity about threefold when measured in the absence of added fructose 6-phosphate and a half-maximal effect was reached at approximately 0.5 mM phosphate. Like glycerol phosphate, phosphate counteracted the inhibition of fructose 2,6-bisphosphatase by fructose 6-phosphate, but a much higher concentration of phosphate than of glycerol phosphate was required to reach this effect.  相似文献   

6.
The 5'-phosphomonoesterase activity of 5'-nucleotidase (EC 3.1.3.5) and alkaline phosphatase (EC 3.1.3.5) participates in the catabolism of purine ribonucleotides to uric acid in humans. Initial velocity studies of 5'-nucleotidase suggest a sequential mechanism of interaction between AMP nad MgCl2, with a Km of 14 and 3 muM, respectively. With product inhibition studies the apparent Ki's for adenosine, inosine, cytidine, and inorganic phosphate were 0.4, 3.0, 5.0, and 42 mM, respectively. A large number of nucleoside mono-, di-, and tri-phosphate compounds were inhibitors of the enzyme. Allopurinol ribonucleotide, ADP, or ATP were competitive inhititors when AMP was the substrate, with a Ki slope of 120 muM. The phosphomonoesterase activity of human placental microsomal alkaline phosphatase had a pH optimum of 10.0 and had only 18% of maximum activity at pH 7.4. Substrates and inhibitors included almost any phosphorylated compound. The Km for AMP was 0.4 mM and the apparent Ki for Pi was 0.6 mM. Activity was increased only 19% by 5 mM MgCl2. These observations suggest that 5'-nucleotidase and alkaline phosphatase may be inhibited by ATP and Pi, respectively, under normal intracellular conditions, and that AMP may be preferentially hydrolyzed by 5'-nucleotidase.  相似文献   

7.
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   

8.
Similarities in properties of ribulose diphosphate carboxylase and oxygenase activities further substantiate the hypothesis that the same protein catalyzes both reactions. The Km (ribulose diphosphate) is 0.33 mM for the ribulose diphosphate oxygenase, when assayed in air with an oxygen electrode. Maximum activity is obtained with 10 to 35 mM MgCl2. Higher MgCl2 concentrations are inhibitory, but they shift the pH optimum from 9.3 or 9.4 to 8.7 or 9.0. MnCl2 is an effective cofactor of the oxygenase and some activity is obtained with CoCl2. Both the ribulose diphosphate carboxylase and oxygenase activity of the purified protein from spinach leaves are slowly inactivated by storage at 0 degrees and reactivated in 10 min at 50 degrees, provided both 25 mM MgCl2 and 1 mM dithiothreitol are present. The sulfhydryl groups of the enzyme which react rapidly with 5,5'-dithiobis(2-nitrobenzoic acid) are approximately 4 at pH 7.8 and 11 at pH 9.4. At both pH values ribulose diphosphate prevents two of these sulfhydryl groups from reacting with this reagent. About 50% inhibition of the oxygenase activity at pH 9.0 occurs with 50 mM bicarbonate in the presence of 3 mM ribulose diphosphate, and from variations in these parameters the inhibition is attributed to the CO2 species. The purified enzyme of acrylamide gels prevented the reduction of nitroblue tetrazolium in the presence of the superoxide radical, but the enzyme in solution did not react as a superoxide dismutase.  相似文献   

9.
Acetohydroxy acid synthase (EC 4.1.3.18) of the archaebacterium Methanococcus aeolicus was purified 1,150-fold to homogeneity. The molecular weight of the purified enzyme was 125,000, and it contained only one type of subunit (M(r) = 58,000). The amino-terminal sequence had 46 to 57% similarity to those of the large subunits of the eubacterial anabolic enzymes and 37 to 43% similarity to those of the yeast and plant enzymes. The methanococcal enzyme had a pH optimum of 7.6. The pI, estimated by chromatofocusing, was 5.6. Activity required Mg2+ or Mn2+ ions, thiamine pyrophosphate, and a flavin. Flavin adenine dinucleotide, flavin mononucleotide, and riboflavin plus 10 mM phosphate all supported activity. However, activity was strongly inhibited by these flavins at 0.3 mM. The Michaelis constants for pyruvate, MgCl2, MnCl2, thiamine pyrophosphate, flavin adenine dinucleotide, and flavin mononucleotide were 6.8 mM, 0.3 mM, 0.16 mM, 1.6 microM, 0.4 microM, and 1.3 microM, respectively. In cell extracts, the enzyme was sensitive to O2 (half-life = 2.7 min with 5% O2 in the headspace), but the purified enzyme was less sensitive to O2 (half-life = 78.0 min with 20% O2). Reconstitution of the enzyme with flavin adenine dinucleotide increased the sensitivity to O2. Moreover, in the assay the homogeneous enzyme was rapidly inactivated by O2, and the concentration required for 50% inhibition (I50) was obtained with an atmosphere of 0.11% O2. The methanococcal enzyme has similarities to the eubacterial and eucaryotic enzymes, consistent with the ancient origin of the archaebacterial enzyme.  相似文献   

10.
The stromal concentration of orthophosphate in intact spinach chloroplasts (prepared in the absence of orthophosphate or pyrophosphate but supplied with both in the reaction medium) fell from a value of approx. 20 mM in the dark to a steady-state concentration of approx. 8 mM in the light. Chloroplasts illuminated in the absence of orthophosphate or pyrophosphate showed a similar trend. However, in this situation the stromal inorganic phosphate (Pi) concentration rapidly decreased from approx. 10 mM in the dark to a constant steady-state concentration of between 1.5 and 2.5 mM in the light. This Pi concentration was not further diminished (even though CO2-dependent O2 evolution had ceased) and was therefore considered to be stromal orthophosphate not freely available to metabolism. In the Pi-deficient chloroplasts the rate of photosynthesis declined rapidly after 1–2 min in the light such that CO2-dependent O2 evolution ceased with 5 min of the onset of illumination. The decline in O2 evolution was accompanied by an increase in the transthylakoid ΔpH (as measured by 9-aminoacridine fluorescence quenching) and in the high-energy state, non-photochemical component of chlorophyll fluorescence quenching (qE). Measurements of stromal metabolite concentrations showed that the ATP/ADP ratio was decreased in the Pi-deficient chloroplasts relative to chloroplasts illuminated in the presence of Pi. The stromal concentration of glycerate 3-phosphate was comparable in the Pi-deficient chloroplasts and those to which Pi had been supplied. Chloroplasts which were illuminated in Pi-free media showed a large accumulation of ribulose-1,5-bisphosphate relative to those supplied with Pi, suggesting inhibition of ribulose-1,5-bisphosphate carboxylase under these conditions. When Pi was added to chloroplasts illuminated in the absence of Pi, both non-photochemical quenching (qE), photochemical quenching (qQ) and ΔpH increased. This suggests that electron transport was not limited by inability to discharge transthylakoid ΔpH. These observation are consistent with the hypothesis that Pi limitation results in decreased ATP production by the thylakoid ATP synthase. The data presented here show that there are multiple sites of flux control exerted by low stromal Pi in the chloroplast. At least three factors contribute to the inhibition of photosynthesis under phosphate limitation: (1) there appears to be a direct effect of Pi on the energy-transducing system; (2) there is direct inhibition of the Calvin cycle decreasing the ability of the pathway to act as a sink for ATP and NADPH; and (3) feedback inhibition of primary processes occurs either via ΔpH or the redox state of electron carriers. However, ΔpH does not appear to be a limiting factor, but rather an inability to regenerate NADP as electron acceptor is suggested. The addition of DCMU to chloroplasts during illumination in the absence of Pi for periods of up to 10 min showed that there was very little loss of variable fluorescence despite a 60% reduction in the capacity for O2 evolution. This would suggest that photoinhibitory damage to Photosystem II was not the major cause of the inhibition of photosynthesis observed with low Pi.  相似文献   

11.
1. Dihydroxyacetone phosphate in concentrations greater than or equal to 2.5 mM completely inhibits CO2-dependent O2 evolution in isolated intact spinach chloroplasts. This inhibition is reversed by the addition of equimolar concentrations of Pi, but not by addition of 3-phosphoglycerate. In the absence of Pi, 3-phosphoglycerate and dihydroxyacetone phosphate, only about 20% of the 14C-labelled intermediates are found in the supernatant, whereas in the presence of each of these substances the percentage of labelled intermediates in the supernatant is increased up to 70-95%. Based on these results the mechanism of the inhibition of O2 evolution by dihydroxyacetone phosphate is discussed with respect to the function of the known phosphate translocator in the envelope of intact chloroplasts. 2. Although O2 evolution is completely suppressed by dihydroxyacetone phosphate, CO2 fixation takes place in air with rates of up to 65 mu mol-mg1 chlorophyll-h1. As non-cyclic electron transport apparently does not occur under these conditions, these rates must be due to endogenous pseudocyclic and/or cyclic photophosphorylation. 3. Under anaerobic conditions, the rates of CO2 fixation in presence of dihydroxyacetone phosphate are low (2.5-7 mumol-mg1 chlorophyll-h1), but they are strongly stimulated by addition of dichlorophenyl-dimethylurea (e.g. 2-10(-7) M) reaching values of up to 60 mumol-mg1 chlorophyll-h1. As under these conditions the ATP necessary for CO2 fixation can be formed by an endogenous cyclic photophosphorylation, the capacity of this process seems to be relatively high, so it might contribute significantly to the energy supply of the chloroplast. As dichlorophenyl-dimethylurea stimulates CO2 fixation in presence of dihydroxyacetone phosphate under anaerobic but not under aerobic conditions, it is concluded t-at only under anaerobic conditions an "overreduction" of the cyclic electron transport system takes place, which is removed by dichlorophenyl-dimethylurea in suitable concentrations. At concentrations above 5-10(-7) M dichlorophenyl-dimethylurea inhibits dihydroxyacetone phosphate-dependent CO2 fixation under anaerobic as well as under aerobic conditions in a similar way as normal CO2 fixation. Therefore, we assume that a properly poised redox state of the electron transport chain is necessary for an optimal occurrence of endogenous cyclic photophosphorylation. 4. The inhibition of dichlorophenyl-dimethylurea-stimulated CO2 fixation in presence of dihydroxyacetone phoshate by dibromothymoquinone under anaerobic conditions indicated that plastoquinone is an indispensible component of the endogenous cyclic electron pathway.  相似文献   

12.
In isolated perfused rat liver, urea synthesis from ammonium ions was dependent on extracellular HCO3- and CO2 concentrations when the HCO3-/CO2 ratio in the influent perfusate was constant (pH 7.4). Urea synthesis was half-maximal at HCO3- = 4 mM, CO2 = 0.19 mM and was maximal at HCO3- and CO2 concentrations above 20 mM and 0.96 mM, respectively. At physiological HCO3- (25 mM) and CO2 (1.2 mM) concentrations in the influent perfusate, acetazolamide, the inhibitor of carbonic anhydrase, inhibited urea synthesis from ammonium ions (1 mM) by 50-60% and led to a 70% decrease in citrulline tissue levels. Acetazolamide concentrations required for maximal inhibition of urea synthesis were 0.01-0.1 mM. At subphysiological HCO3- and CO2 concentrations, inhibition of urea synthesis by acetazolamide was increased up to 90%. Inhibition of urea synthesis by acetazolamide was fully overcome in the presence of unphysiologically high HCO3- and CO2 concentrations, indicating that the inhibitory effect of acetazolamide is due to an inhibition of carbonic-anhydrase-catalyzed HCO3- supply for carbamoyl-phosphate synthetase, which can be bypassed when the uncatalyzed intramitochondrial HCO3- formation from portal CO2 is stimulated in the presence of high portal CO2 concentrations. With respect to HCO3- supply of mitochondrial carbamoyl-phosphate synthetase, urea synthesis can be separated into a carbonic-anhydrase-dependent (sensitive to acetazolamide at 0.5 mM) and a carbonic-anhydrase-independent (insensitive to acetazolamide) portion. Carbonic-anhydrase-independent urea synthesis linearly increased with the portal 'total CO2 addition' (which was experimentally determined to be CO2 addition plus 0.036 HCO3- addition) and was independent of the perfusate pH. At a constant 'total CO2 addition', carbonic-anhydrase-dependent urea synthesis was strongly affected by perfusate pH and increased about threefold when the perfusate pH was raised from 6.9 to 7.8. It is concluded that the pH dependent regulation of urea synthesis is predominantly due to mitochondrial carbonic anhydrase-catalyzed HCO3- supply for carbamoyl phosphate synthesis, whereas there is no control of urea synthesis by pH at the level of the five enzymes of the urea cycle. Because HCO3- provision for carbamoyl phosphate synthetase increases with increasing portal CO2 concentrations even in the absence of carbonic anhydrase activity, susceptibility of ureogenesis to pH decreases with increasing portal CO2 concentrations. This may explain the different response of urea synthesis to chronic metabolic and chronic respiratory acidosis in vivo.  相似文献   

13.
1. A method is described for the purification of a form of 3-deoxy-D-arabinoheptulosonate-7-phosphate synthase (tryptophan) that probably differs from that of the native enzyme. 2. The kinetics of the reaction catalysed by 3-deoxy-D-arabinoheptulosonate-7-phosphate synthase (tryptophan) shows that the reaction proceeds via a ping-pong bi-bi mechanism, with activation by phosphoenolpyruvate (P-Prv), the first substrate, and inhibition by erythrose 4-phosphate (Ery-P) the second substrate. At low substrate concentrations, KP-Prv is 0.1 mM and KEry-P is 0.13 mM. 3. The substrates phosphoenolpyruvate and erythrose 4-phosphate and the product inorganic phosphate can protect the purified enzyme against heat denaturation, whereas the inhibitor, tryptophan, has no effect, although it binds to the enzyme in the absence of other ligands. 4. Product inhibition by inorganic phosphate is linear non-competitive with respect to phosphoenolpyruvate (Ki, slope = 22 mM and Ki, intercept = 54 mM) and substrate-linear competitive with respect to erythrose 4-phosphate (Ki, slope = 25 mM). 5. The enzyme has an activity optimum at pH 7.3 and a tryptophan inhibition optimum at pH 6.4, Trp 0.5 is 4 microM. Inhibition by tryptophan is non-competitive with respect to phosphoenolpyrovate and substrate-parabolic competitive with respect to erythrose 4-phosphate. 6. The role of the enzyme in metabolic regulation is discussed.  相似文献   

14.
Some kinetic properties of N-acetylglutamate 5-phosphotransferase (ATP: N-acetyl-L-glutamate 5-phosphotransferase EC 2.7.2.8) purified approx. 2000-fold from Pseudomonas aeruginosa have been studied. The enzyme required Mg2+ for activity. Mn2+, Zn2+, Co2+, and Ca2+, in this order, could replace Mg2+ partially. The substrate specificity was narrow: N-carbamoyl-L-glutamate and N-formyl-L-glutamate were phosphorylated, but at a lower rate than N-acetyl-L-glutamate; N-propionyl-L-glutamate was almost inactive as a substrate. dATP, but neither GTP nor ITP, could be used instead of ATP. The enzyme had a broad pH optimum from pH 6.5 to 9. Feedback inhibition by L-arginine was markedly dependent on pH. Above pH 9 no inhibition was observed. L-Citrulline was three times less potent an inhibitor than L-arginine. The enzyme showed Michaelis-Menten kinetics, even at low concentration of the second substrate. The apparent Km was 2 mM for N-acetyl-L-glutamate (at 10 mM ATP) and approx. 3 mM for ATP (at 40 mM N-acetyl-L-glutamate). In the presence of L-arginine the rate-concentration curves for N-acetyl-L-glutamate became signoidal, while no cooperativity was detected for ATP. A method was developed allowing the determination of N-acetyl-L-glutamate in the nanomolar range by means of purified enzyme.  相似文献   

15.
The present work compares the effects of several ligands (phosphatase substrates, MgCl2, RbCl and inorganic phosphate) and temperature on the phosphatase activity and the E2(Rb) occluded conformation of Na+/K+-ATPase. Cooling from 37 degrees C to 20 degrees C and 0 degrees C (hydrolysis experiments) or from 20 degrees C to 0 degrees C (occlusion experiments) had the following consequences: (i) dramatically reduced the Vmax for p-nitrophenyl phosphate and acetyl phosphate hydrolysis but it produced little or no changes in the Km for the substrates; (ii) led to a 5-fold drop in the Km for the inorganic phosphate-induced di-occlusion of E2(Rb); (iii) reduced the K0.5 and curve sigmoidicity of the Rb-stimulated hydrolysis of p-nitrophenyl phosphate and acetyl phosphate and the Rb-promoted E2(Rb) formation. At 20 degrees C, in the presence of 1 mM RbCl and no Mg2+, acetyl phosphate did not affect E2(Rb); with 3 mM MgCl2, acetyl phosphate stimulated a release of Rb from E2(Rb) both in the presence and absence of RbCl in the incubation mixture. As a function of acetyl phosphate concentration the Km for iRb release was indistinguishable from the Km found for stimulation of hydrolysis and enzyme phosphorylation under identical experimental conditions; in addition, the extrapolated di-occluded fraction corresponding to maximal hydrolysis was not different from 100%. These results indicate that although E2(K) might be an intermediary in the phosphatase reaction, the most abundant enzyme conformation during phosphatase turnover is E2 which has no K+ occluded in it. The ligand interactions associated to phosphatase activity do not support an equivalence of this reaction with the dephosphorylation step in the Na+ + K+-dependent ATP hydrolysis; on the other hand, there are similarities with the reversible binding of inorganic phosphate in the presence of Mg2+ and K+ ions.  相似文献   

16.
Human erythrocytes contain a phosphatase that is highly specific for phosphoglycollate. It shows optimum pH of 6.7 and has Km 1 mM for phosphoglycollate. The molecular weight appears to be about 72000. The enzyme is a dimeric molecule having subunits of mol. wt. about 35000. It could be purified approx. 4000-fold up to a specific activity of 5.98 units/mg of protein. The activity of the enzyme is Mg2+-dependent. Co2+, and to a smaller extent Mn2+, may substitute for Mg2+. Half-maximum inhibition of the phosphatase by 5,5'-dithiobis-(2-nitrobenzoate), EDTA and NaF is obtained at 0.5 microM, 1 mM and 4 mM respectively. Moreover, it needs a univalent cation for optimum activity. Phosphoglycollate phosphatase is a cytoplasmic enzyme. Approx. 5% of its total activity is membrane-associated. This part of activity can be approx. 70% solubilized by freezing, thawing and treatment with 0.25% Triton X-100.  相似文献   

17.
When the R-form lipopolysaccharide (LPS) from Klebsiella pneumoniae strain LEN-111 (O3-:K1-), from which cationic material had been removed by electrodialysis, was suspended in 50 mM Tris buffer at pH 8.5 containing 0.1 mM or higher concentrations of MgCl2, it formed an ordered two-dimensional hexagonal lattice structure and its center-to-center distance (lattice constant) depended upon the concentration of MgCl2 and reached the shortest value (14 nm) at 10 mM. In contrast, in the presence of 0.1 to 10 mM CaCl2 in place of MgCl2, the electrodialyzed LPS did not form such an ordered hexagonal lattice structure but formed an irregular network structure with a center-to-center distance of 19 to 20 nm. We investigated interaction of Mg2+ and Ca2+ in formation of the hexagonal lattice structure by the electrodialyzed LPS suspended in 50 mM Tris buffer at pH 8.5. When 0.1 mM or higher concentrations of CaCl2 were mixed with 1 mM MgCl2 or when 1 mM or higher concentrations of CaCl2 was mixed with 10 mM MgCl2, the electrodialyzed LPS did not form the hexagonal lattice structure of the magnesium salt type but formed the irregular network structure of the calcium salt type. In the coexistence of equimolar or higher concentrations of CaCl2 together with 1 or 10 mM MgCl2, the binding of Mg to the electrodialyzed LPS was significantly inhibited and, conversely, the binding of Ca was enhanced as compared with when MgCl2 or CaCl2 was present alone. However, the coexistence of 10 times less molar concentrations of CaCl2 did not significantly inhibit the binding of Mg to the electrodialyzed LPS. Therefore, the inhibition of formation of the Mg2(+)-mediated hexagonal lattice structure of the electrodialyzed LPS by equimolar or higher concentrations of CaCl2 accompanied the inhibition of binding of Mg but that by 10 times less molar concentrations of CaCl2 did not accompany it.  相似文献   

18.
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   

19.
The intrinsic fluorescence of sarcoplasmic reticulum vesicles was measured under conditions allowing ATPase phosphorylation from inorganic phosphate. Significant fluorescence enhancement of up to 4% resulted from gradient-independent enzyme phosphorylation at pH 6, in the absence of KCl. The equilibrium fluorescence data obtained at various magnesium and phosphate concentrations agree with a reaction scheme in which Mg2+, as direct activator, and free phosphate, as the true substrate, bind to the enzyme in random order to give a noncovalent ternary complex (Mg.*E.Pi), in equilibrium with the covalent phosphoenzyme (Mg.*E-P). The transient kinetics of the fluorescence rise was also studied, and the resulting data were generally consistent with the above scheme, assuming that binding reactions are fast compared to covalent phosphoenzyme formation. This, however, might be valid only as a first approximation. At 20 degrees C and pH 6, the phosphate concentration for half-maximum phosphorylation rate constant, at 20 mM magnesium, was higher than 20 mM. Similarly, the magnesium concentration for half-maximum phosphorylation rate constant, at 20 mM phosphate, was also higher than 20 mM. The maximum phosphorylation rate was faster than 25 s-1, and the phosphoenzyme hydrolysis rate constant was 1.5-2 s-1 under these conditions, so that the equilibrium constant between Mg.*E.Pi and Mg.*E-P largely favors the phosphoenzyme.  相似文献   

20.
Retinylmonophosphatase (RMPase) activity in mouse brain paralleled the subcellular distribution of the plasma-membrane marker Na+ + K+-dependent ATPase. The enzyme had a pH optimum between 5.5 and 7.0. The enzyme demonstrated linear kinetics with respect to time and both protein and substrate concentrations. RMPase was saturated by low retinyl monophosphate (RMP) concentrations and exhibited an apparent Km of 4.6 microM. The enzyme did not require MgCl2 for activity, and in fact assays were routinely run in the presence of 10 mM-Na2EDTA. In general, detergents inhibited the enzyme, with 0.05% Triton X-100 causing a 30% loss of activity. Phosphatidic acid was also inhibitory, but phosphatidylcholine and sphingomyelin stimulated phosphatase activity. RMPase was inhibited 35% by 5 mM concentrations of fluoride, phosphate or pyrophosphate. A series of other phosphorylated compounds, including glucose 6-phosphate, alpha-glycerophosphate, ATP, AMP, p-nitrophenyl phosphate and thiamin pyrophosphate, showed little or no inhibition. RMPase activity differed in several characteristics from that previously reported for dolichylmonophosphatase. It is concluded that RMP could play a distinct role in the plasma membrane.  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