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.
The relationship between area per molecule and surface pressure of monolayers of phosphatidylglycerol phosphate from extreme halophile Halobacterium cutrirubrum and its deoxy analogue, deoxyphosphatidylglycerol phosphate, spread at an air/water interface was examined. The effect of ionization of the primary and secondary acidic functions of the phosphate groups of the two lipids on surface characteristics of compression isotherms was determined by spreading monolayers on subphases with pH values ranging from below the apparent pKa of the primary ionization (pH 0) to greater than that of secondary ionization (pH 10.9). The limiting molecular area increases with decreasing pH below 2. Ionization of the primary phosphate functions of both phospholipids (with bulk pK1 values close to 4) is associated with a marked expansion of the films, as judged by values of limiting molecular area. Ionization of the secondary phosphate functions causes further expansion of the films, with the apparent pK2 of deoxyphosphatidylglycerol phosphate slightly less than that indicated for phosphatidylglycerol phosphate. Values of surface-compressibility modulus calculated from the surface characteristics of the phosphatidylglcerol phosphate monolayers showed that films spread on subphases with a pH of about the apparent pK1 of the primary phosphate functions were the least compressible. Increasing or decreasing subphase pH caused an increase in compressibility; this effect on compressibility was much less with monolayers of deoxyphosphatidylglycerol phosphate at high pH. The effect of inorganic counter-ions on monolayer characteristics of phosphatidylglycerol phosphate was examined by using subphases of NaCl concentrations varying from 0.01 to 1 M. The limiting molecular area was found to increase exponentially with respect to the subphase NaCl concentration.  相似文献   

2.
Bovine beta 2-microglobulin (beta 2-m), the light chain of the histocompatibility antigen, was isolated in crystalline form from colostrum. Previous studies from this laboratory on the solution properties of this protein suggest the existence of a time-dependent multiple aggregation phenomenon. To clarify the molecular states of beta 2-m, its solution properties were studied by ultracentrifugation and spectropolarimetry. Sedimentation equilibrium experiments at pH 5.0 (0.08 M NaCl, 0.02 M sodium phosphate) at concentrations less than 0.3 mg/ml give Mr = 11,800. From sedimentation velocity results, we conclude that bovine beta 2-m is a much more symmetrical and compact molecule than either guinea pig or human beta 2-m. At concentrations above 0.4 mg/ml under the same conditions, sedimentation equilibrium experiments show that a monomer to tetramer reversible self-association occurs. Also, the tetramerization increases with decreasing temperature. beta 2-Microglobulin undergoes an irreversible temperature-dependent association to a much larger aggregate over a period of 7 days, as evidenced by sedimentation equilibrium and velocity results. The rate of this aggregation decreases as the pH approaches the isoelectric point (pH 7) from either side. Furthermore, circular dichroism measured at pH 5.0 under time-dependent aggregating conditions showed a marked increase in the percentage of disordered structure, leading to the conclusion that this effect is a denaturation phenomenon.  相似文献   

3.
The inactivation kinetics of Listeria monocytogenes in a phosphate buffer (PB) was determined at different hypochlorite concentrations, pH values and temperatures. D-values, using a linear regression, of L. monocytogenes in PB (pH 6.5) were 23.54, 17.40, 14.24 and 12.00s at 5, 10, 50 and 100 mg l(-1) hypochlorite, respectively, at 30 degrees C. The k-values ranged from 0.098 to 0.192s(-1) and 0.007 to 0.018s(-1) for hypochlorite concentrations (from 5 to 100 mg l(-1)) in PB (pH 6.5) and PB containing 0.1% peptone (pH 6.5), respectively, at 30 degrees C. D-values of L. monocytogenes exposed to hypochlorite were decreased with decreasing pH of PB (pH from 8.5 to 4.5). Hypochlorite showed higher antimicrobial activity at higher temperature. Not only the effect of hypochlorite concentration on the inactivation of L. monocytogenes but also other parameters like temperature, pH and suspending solutions effect the inactivation rates.  相似文献   

4.
Ornithine carbamoyltransferase (OCT) activity was detected in extracts from mature leaves, fruit, germinating seeds, and seedlings of Vitis vinifera L. Michaelis-Menten constants for OCT were 3.5 millimolar for carbamyl phosphate and 5.5 millimolar for l-ornithine. Concentrations of l-ornithine greater than 10 millimolar slightly inhibited the enzyme, whereas carbamyl phosphate at concentrations greater than the optimal (about 10 millimolar) did not affect OCT activity. l-Citrulline formation was linear with incubation period for the first 25 minutes and with increasing amounts of enzyme up to an equivalent of about 200 milligrams of fresh tissue. The optimum pH for in vitro OCT activity was between 8.4 and 8.8, and the optimum incubation temperature was 38 C.  相似文献   

5.
Acid phosphatase-1 (orthophosphoric monoester phosphohydrolase, acid optimum, EC 3.1.3.2), the major phosphatase in adult Drosophila melanogaster, has been purified to apparent homogeneity. The final product is a glycoprotein homodimer with a subunit molecular weight of about 50,000, as measured by its electrophoretic mobility in denaturing conditions on polyacrylamide gels containing sodium dodecyl sulfate. It has a turnover number of 1720 1-naphthyl phosphate molecules hydrolyzed/s by each acid phosphatase-1 molecule at 37 degrees C, pH 5.0. An average fly contains about 5 ng of enzyme. Pure acid phosphatase-1 displays heterogeneity in isoelectric focusing, with a major band at pH 5.3. The enzyme hydrolyzes a wide variety of phosphate monoesters, including AMP, glucose 6-phosphate, ATP, choline phosphate, or phosphoproteins. The maximum reaction rates are different for all substrates, and some substrates appear to inhibit the reaction at high substrate concentrations. The Michaelis constants for 1-naphthyl phosphate and p-nitrophenyl phosphate are 79 microM and 68 microM, respectively, at pH 5.0 and 37 degrees C. The optimum pH level for 1-naphthyl phosphate is 4.5. Acid phosphatase-1 is inhibited by L(+)-tartrate (but not D(-)-tartrate), phosphate, and fluoride. The reaction rate increases 2.1-fold for every 10 degrees C rise in temperature. Above 48 degrees C, the rate of thermal denaturation is greater than the rate of the enzyme reaction.  相似文献   

6.
The adsorption of phosphate on metal (hydr)oxides may be influenced by the pH and by the adsorption of other ions. In this study, the influence of sulphate and pH on phosphate adsorption on goethite and the availability to plants of adsorbed phosphate was examined. Maize plants were grown on suspensions of goethite with adsorbed phosphate, containing the same total amount of phosphate and either 0.11 mM or 2.01 mM sulphate at pH 3.7, 4.6 or 5.5. The uptake of phosphorus by the plants increased with the larger sulphate concentration and decreasing pH. Mean P uptake in the treatment with 2.01 mM sulphate and pH 3.7 was 55 µmol plant-1, whereas in the treatment with 0.11 mM sulphate and pH 5.5 it was 2 µmol plant-1. Batch adsorption experiments using32 P and speciation modelling of ion adsorption showed that in the presence of sulphate, the phosphate concentration in solution strongly increased with decreasing pH, due to competitive adsorption between sulphate and phosphate on goethite. Modelled phosphate concentrations in solution in the uptake experiment were all below 0.6 µM and correlated well with the observed P uptake. This correlation indicates that the strong influence of the sulphate concentration and pH on the plant-availability of adsorbed phosphate results from the competition between sulphate and phosphate for adsorption on goethite.  相似文献   

7.
The stability of prostaglandin E1 (PGE1) in three physiologic solutions was studied at body temperature (37 degrees C) over 32 days. The solutions were 100 mcg/ml PGE1 in isotonic saline (pH 4.5), 0.1 M phosphate buffered water (pH 7.4) or 0.01 M phosphate buffered isotonic saline (pH 4.7). PGE1 was found to be more stable in the saline and buffered saline solutions at the pH values of 4.5 and 4.7 respectively. Twenty-five per cent of the PGE1 remained at 32 days in these solutions while 95% of the PGE1 in the solution at pH 7.4 was degraded by day 14. The degradation of PGE1 in the acidic solutions appeared to be nearly linear when plotted on a semilog graph. This data allows one to use PGE1 in an aqueous, slightly acidic solution in a system that requires it to be kept at 37 degrees C for up to 30 days such as a biologically implantable pump. Investigators can use such a system in vivo to study the effect of known concentrations of PGE1 given over a period of time to a specific area of interest.  相似文献   

8.
Steven C. Huber 《Planta》1980,149(5):485-492
Chloroplast stromal volume and pH influenced the phosphate (Pi)-dependence of photosynthesis of spinach (Spinacia oleracea L.) chloroplasts. Decreasing the sorbitol concentration in the reaction mixture from 0.35 to 0.25 M, or decreasing the external pH from 8.3 to 7.3, extended the induction period of photosynthesis and decreased both the optimal [Pi] and the minimal [Pi] required to inhibit O2 evolution completely. At least part of the effect of external pH was attributable to changes in stromal pH on the basis of effects of NH4Cl and sodium acetate at a constant external pH. When the external pH was increased from 7.3 to 8.3, the stromal pH changed only about 0.6 pH units. Hence, the pH gradient across the envelope was diminished and the efflux of phosphoglycerate relative to dihydroxyacetone phosphate was enhanced.Calvin-cycle activity, varied with light intesity or electron transport inhibitors, affected the rate of photosynthesis but not the induction period or the Pi optimum for photosynthesis. Relatively low Calvin-cycle activity was apparently sufficient to fill metabolite pools and thus terminate the induction period. The results indicate that pH does not affect the Pi dependence of photosynthesis by reducing Calvin-cycle activity. Rather, it is postulated that at low stromal pH, larger metabolic pools are required to maintain maximum rates of photosynthesis because of changes in substrate affinity of some Calvin-cycle enzymes. Consequently, chloroplast photosynthesis would be more sensitive to exogenous Pi.Abbreviations DHAP dihydroxyacetone phosphate - PGA 3-phosphoglycerate - Pi inorganic phosphate Cooperative investigations of the North Carolina Agricultural Research Service and Agricultural Research, Science and Education Administration, U.S. Department of Agriculture, Raleigh, N.C. Paper No. 6391 of the Journal Series of the North Carolina Agricultural Research Service, Raleigh, N.C., USA  相似文献   

9.
Ornithine carbamoyltransferase (OCT) activity was detected in extracts from mature leaves, fruit, germinating seeds, and seedlings of Vitis vinifera L. Michaelis-Menten constants for OCT were 3.5 millimolar for carbamyl phosphate and 5.5 millimolar for l-ornithine. Concentrations of l-ornithine greater than 10 millimolar slightly inhibited the enzyme, whereas carbamyl phosphate at concentrations greater than the optimal (about 10 millimolar) did not affect OCT activity. l-Citrulline formation was linear with incubation period for the first 25 minutes and with increasing amounts of enzyme up to an equivalent of about 200 milligrams of fresh tissue. The optimum pH for in vitro OCT activity was between 8.4 and 8.8, and the optimum incubation temperature was 38 C.  相似文献   

10.
The effect of pH in the range 5.0 to 7.0 on the thermal destruction of spores of Clostridium sporogenes putrefactive anaerobe 3679 was examined by three methods: a capillary tube method in which spores were suspended in phosphate buffers, a thermoresistometer method in which spores were suspended in buffered pea puree adjusted to the same set of pH values, and a thermal death time can method in which spores were again suspended in buffered pea puree. The results indicated that increasing acidity is, in general, accompanied by decreasing heat resistance, although the pH effect was more pronounced at the higher than at the lower processing temperatures. Certain pH values appear to be critical, as they produced, in all three sets of experiments, effects which would not be predicted by the overall relationship between acidity and spore heat resistance. Differences between heat resistance in phosphate buffer as compared with that in pea puree adjusted to the same pH were also noted. D-values in buffer were found to be lower than those in pea puree, except at the highest temperatures coupled with the lowest pH values. The differences between buffer D-value and pea puree D-value were found to increase with increasing pH and with decreasing temperature. On the other hand, at all pH values examined, z-values determined in buffer were somewhat higher than those determined in pea puree adjusted to the same pH.  相似文献   

11.
The pH change of 10 different buffering systems with temperature ranging from room to 5 °C was examined; three were conventional buffers which included phosphate yolk, citrate yolk, and skim milk. Seven were Good's buffers with egg yolk which included TES, TRIS, BES, MOPS, PIPES, MES, and TEST. The pH of the three conventional buffers did not change with decreasing temperature, but Good's buffers showed an increase in pH with decreasing temperature from room to 5 °C. The pH change due to temperature was measured for TEST buffer solution with and without 20% egg yolk containing 2 or 6% of five different cryoprotective compounds. The pH at 5 °C was significantly higher than at room temperature. The addition of egg yolk and/or cryoprotective compound did not alter the pH significantly during cooling, even though a slight drop in pH was noted with the addition of egg yolk indicating that the change in pH is primarily due to the buffer. The pH of TEST yolk buffer (pH 7.2 at room temperature) was measured continuously from 37 °C to below freezing (?18 °C). The pH increased with decreasing temperature to 8.0 ± 0.2 from 37 to ?14 °C at which point it dropped abruptly to pH 6.5 ± 0.2.  相似文献   

12.
31P NMR was applied to an examination of the freeze-tolerant larvae of the gall fly, Eurosta solidaginis. Resonances from sugar phosphates, inorganic phosphate, adenylates and arginine phosphate were identified. Two peaks of Pi were identified corresponding to intracellular and extracellular Pi. Anoxia produced an expected decrease in peak intensities of ATP and arginine phosphate while the peak of intracellular Pi was enhanced and shifted to indicate intracellular acidification during anoxia. Spectra of whole larvae were monitored over a temperature range from -30 degrees to +25 degrees C. No abrupt alterations in the spectra were seen at the point of extracellular freezing which occurs at about -8 degrees C but temperature had dramatic effects upon the peak intensities of ATP and arginine phosphate. A reversible increase/decrease in peak intensities, relative to Pi, was observed as temperature was raised/lowered. At 15 degrees and -20 degrees C, the beta peak of ATP was 64% and 2% of the peak intensity of Pi while that of arginine phosphate was 78% and 11%, respectively. This temperature effect was not an artifact of instrumentation (as model solutions containing Pi, ATP and arginine phosphate did not show this effect) or a result of changes in the total amounts of these compounds in the cell with temperature. Rather it is apparent that these molecules become restricted in their rotational movement as temperature is lowered perhaps via binding to subcellular components. Changes in the amounts of freely soluble ATP and arginine phosphate with temperature could have important implications for metabolism and its control. Analysis of the effect of temperature on the chemical shift of Pi was also used to determine pH in the intracellular and extracellular compartments. Temperature change had no effect on extracellular (hemolymph) pH which remained constant at 6.1-6.3. Intracellular pH varied with temperature, however, from pH 6.8 at 15 degrees C to pH 7.3 at -12 degrees C with a change, delta pH/delta 0, of -0.0185 degrees C consistent with alphastat regulation.  相似文献   

13.
The effects of the neutral salt concentration, pH, and coexistence of myosin on the denaturation of F-actin without ATP at low temperature were studied using the DNase I inhibition assay. The percent denaturation of F-actin gradually increased with a decrease in pH from 8.0 to 5.2, on incubation for 2 weeks in the presence of 50 mM KCl at 0 degrees C. This change was much faster in 0.5 M KCl and more than 75% of the F-actin became denatured on incubation for 1 week at pH 5.2. The buffer composition was found to exert a strong influence on the denaturation of F-actin. That is, there was a tendency for the denaturation of F-actin at pH 6.0 to be faster in MES[2-(N-morpholino)ethanesulfonic acid]-NaOH buffer than in sodium phosphate buffer, the critical concentrations of actin in 0.5 M KCl being 0.31 mg/ml for MES-NaOH buffer and 0.15 mg/ml for sodium phosphate buffer. A sigmoidal relationship was found between the percent denaturation of F-actin and the KCl concentration added, the greatest change occurring at KCl concentrations between 0.25 and 0.75 M. The time courses of the denaturation of F-actin showed that the percent denaturation rose at first and that in time the rate of the increase decreased. In the case of pH 8.0 and 0.5 M KCl, it took about 1 week for the denaturation rate to begin to drop. The pH of 6.0 further promoted the instability of F-actin exposed to high KCl concentrations.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

14.
The kinetics of sodium gradient dependent phosphate uptake by the renal brush border membrane vesicles of the rat have ben studied under various conditions of temperature and pH. From 7 to 30 degrees C the Lineweaver-Burk plots are linear, and the apparent Km progressively increases from 54 to 91 microM. Above 30 degrees C, the apparent Km continues to increase to reach 135 microM at 40 degrees C, but a break is observed in the Lineweaver-Burk plots at the substrate concentration of 300 microM. The existence of this break, confirmed by the Eadie-Hofstee plot supports the hypothesis of a dual mechanism of phosphate transport, one for low concentrations of substrate with a Km of 100 microM and the other for high concentrations with a Km of approximately 240 microM. When the two components of the Eadie-Hofstee plot are analyzed according to a nonlinear regression program, these two values of Km become 70 microM and 1.18 mM, respectively. The Vmax continuously increases with temperature. However, the Arrhenius plot (In Vmax vs. 1/TK) shows an abrupt discontinuity at 23 degrees C. pH experiments were performed at 35 degrees C. In the absence of a proton gradient, increasing the pH from 6.5 to 7.5 and 8.5 decreases the apparent Km from 341 to 167 and 94 microM, respectively. When only the divalent form of phosphate is considered as the substrate, the apparent Km does not vary anymore with the pH and remains around the mean value of 105 microM. The uniformity of the apparent Km for the total phosphate uptake, when only the divalent phosphate is considered as being the substrate, suggests that this divalent form is the only one which is transported. Whatever the substrate considered, total phosphate or divalent phosphate, the highest Vmax is obtained at pH 7.5 which probably approximates the optimum pH inside the vesicles for the phosphate uptake.  相似文献   

15.
The effects of low temperature assay (5 °C) on the properties of the aerobic (low phosphate) vs. anoxic (high phosphate) forms of pyruvate kinase (PK) from foot muscle and gill of the whelk Busycon canaliculatum (L.) were assessed at two pH values, pH 7.00 and 7.25, and compared to control conditions of 20 °C and pH 7.00 (all assayed in imidazole buffer). When pH was held constant at 7.00, the decrease in assay temperature to 5 °C had large effects on the measured kinetic parameters of all PK forms, as compared to 20 °C and pH 7.00. However, when assay pH was allowed to rise, from 7.00 to 7.25, with the temperature decrease to 5 °C there were fewer alterations of kinetic parameters and quantitatively smaller changes to enzyme properties. It appears, then, that when pH rises with decreasing temperature following alphastat predictions, kinetic properties of PK are largely conserved. Low temperature, at either pH value, had several significant effects on PK properties. For example, low temperature raised the S0.5 for phosphoenolpyruvate of PK-anoxic from gill by 3–6 fold and decreased the I50 Mg · ATP for PK-anoxic from foot by the same amount. Arrhenius plots of PK activity for the gill PK forms showed a distinct break at 10 °C; > 10 °C Q10 was 2.5 whereas < 10 °C Q10 was 8.4. Temperature-dependent changes in all cases affected enzyme properties in a manner that would restrict enzyme function at low temperature.  相似文献   

16.
Phosphoenolpyruvate carboxylase from Amaranthus viridis leaves was activated by inorganic orthophosphate in a concentration- and pH-dependent manner. Maximal activation at pH 7.0 was achieved at phosphate concentrations above 20 mM, and a positive cooperativity was observed for the binding of the anion at this pH. At pH 8.0 the maximum of activity was achieved at 10 mM phosphate; higher concentrations reduced the activation. KM for phosphoenolpyruvate-Mg at pH 7.0 was lowered by phosphate in all concentrations tested up to 30 mM. While at pH 8.0 the KM values were lower than that of the control up to 10 mM phosphate; higher anion concentrations raised the minimum value of KM at this pH. VMAX increased at pH 7.0, and remained unchanged at pH 8.0. A KA value of 0.41 mM was calculated for phosphate at the alkaline pH. The phosphate analogue arsenate also behaved as an activating agent, while other anions (e.g. nitrate, nitrite, sulfate, tetraborate) were ineffective. The phosphate-activated enzyme was shown to be insensitive to glucose-6-phosphate, but was inhibited by l -malate to the same extent as the control.  相似文献   

17.
Apotryptophanase (L-tryptophan indole-lyase, EC 4.1.99.1) from Escherichia coli B/1t7A shows, in the presence of potassium phosphate, a temperature-dependent structural rearrangement which is not observed in the presence of sodium phosphate or imidazole plus KC1. This rearrangement can be described by a two-state equilibrium between two forms of the apoenzyme. The midpoint temperature of the rearrangement (TM) and the van't Hoff enthalpy (delta H) at different potassium phosphate concentrations and pH values, respectively, were determined by measuring the temperature-dependence of the ultraviolet absorbance of apotryptophanase. Increasing the potassium phosphate concentration at pH 7.8 causes a simultaneous increase in total absorbance and the delta H value, whereas the TM increases between pH 7.0 and 7.8 but starts to decrease at pH values above 7.8. In 0.1 M potassium phosphate at the pH optimum of the enzyme (7.8) TM and delta H were found to be 293.1 K and 167 kJ X mol-1, respectively. Moreover, the tyrosine residues of apotryptophanase dissociate in potassium phosphate and in imidazole plus KCl with pK values of 8.6 and 9.8, respectively, indicating that potassium phosphate favors the formation of tyrosinate. The rearrangement might be interpreted as the formation of specific hydrogen bonds between tyrosine and potassium phosphate which are ruptured at higher temperature. Such hydrogen bonds cannot be formed at all or only to a small extent in the presence of imidazole plus KCl or sodium phosphate. Those hydrogen bonds stabilize the structure of apotryptophanase. In contrast, holotryptophanase requires only K+ for enzymatic activity.  相似文献   

18.
In order to obtain basic information on ubiquinone (UQ) formation by BY-2 cells in suspension culture, effects of inorganic phosphate and nitrogen sources and such physical factors as initial pH, light irradiation, shaking condition and temperature were investigated. The concentration of phosphate and nitrogen sources had no marked effect, but the ratio of ammonium nitrogen to nitrate nitrogen was significantly effective on UQ formation. The UQ content in BY-2 cells tends to increase at higher ratios of ammonium nitrogen. The increase in the UQ content was recognized at higher concentrations of 2, 4-d, especially with lower concentrations of sucrose. Physical factors had no marked effect on UQ formation except temperature. The UQ content was considerably raised at higher temperatures.  相似文献   

19.
The activity of phospholipase A on phosphatidyl choline and phosphatidal choline spread as monolayers on phosphate buffers containing snake venom (Crotalus atrox or Naja naja) was studied by measuring the fall of surface potential as a function of time, pH, film pressure, temperature, and concentrations of phosphate and venom. At 25 degrees C, pH 7.0, and 0.2 micrograms of venom per ml, optimal activity was observed with both venoms on both substrates at 12 dynes/cm film pressure on 0.04 m phosphate. Under these conditions, the pH optimum for C. atrox was broad (6.6-7.4) and that for N. naja was sharp (8.0) for the action on phosphatidyl choline, whereas both venoms had a sharp optimum at pH 8.0 in their action on phosphatidal choline. The optimal temperature with phosphatidyl choline was 27.5 degrees C for N. naja and 40 degrees C for C. atrox. In line with studies of phospholipase A activity in bulk phase in ether, phosphatidal choline was attacked much more slowly than phosphatidyl choline by C. atrox. Under conditions where both venoms had equal activity on phosphatidyl choline, C. atrox was only half as active as N. naja on phosphatidal choline. The studies suggest that the linkage of the hydrophobic chains in glycerophosphatides may affect their interaction with proteins.  相似文献   

20.
We re-examined the kinetics of the bisphosphatase reaction of rat hepatic 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatase after depleting the enzyme of bound fructose 6-phosphate and found a hyperbolic dependence on fructose 2,6-bisphosphate at concentrations below 100 nM. The Michaelis constant was 4 nM, the Vmax was about 12 nmol X mg-1 X min-1 at 22 degrees C but the substrate inhibited at concentrations above 100 nM. Both phosphate and alpha-glycerol phosphate strongly inhibited phosphoenzyme formation and hydrolytic rate below 100 nM, but relieved the inhibition by substrate at higher concentrations probably by antagonizing substrate binding. A number of observations support the proposition that the phosphoenzyme is a necessary participant in catalysis. 1) The amount of phosphoenzyme measured during steady-state hydrolysis as a function of substrate concentration correlated with the velocity profile. 2) Rapid mixing experiments demonstrated that over a broad range of substrate concentrations phosphoenzyme formation was faster than the net rate of hydrolysis. 3) Both phosphate and alpha-glycerol phosphate inhibited the rate of phosphoenzyme formation and, at low substrate concentrations, reduced the steady-state phosphoenzyme levels. The latter correlated with inhibition of substrate hydrolysis. 4) Both phosphate and alpha-glycerol phosphate stimulate the rate of phosphoenzyme breakdown, consistent with their stimulation of substrate hydrolysis at high substrate concentrations. 5) The fractional rate of phosphoenzyme breakdown, which was pH and substrate dependent, multiplied by the amount of phosphoenzyme obtained in the steady state at that pH and substrate concentration approximated the observed rate of hydrolysis. We conclude that the phosphoenzyme is a reaction intermediate in the hepatic fructose-2,6-bisphosphatase reaction.  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