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.
2.
1. The pH in the stroma and in the thylakoid space has been measured in a number of chloroplast preparations in the dark and in the light at 20 degrees C. Illumination causes a decrease of the pH in the thylakoid space by 1.5 and an increase of the pH in the stroma by almost 1 pH unit. 2. CO2 fixation is shown to be strongly dependent on the pH in the stroma. The pH optimum was 8.1, with almost zero activity below pH 7.3.Phosphoglycerate reduction, which is a partial reaction of CO2 fixation, shows very little pH dependency. 3. Low concentrations of the uncoupler m-chlorocarbonylcyanide phenylhydrazone (CCCP) inhibit CO2 fixation without affecting phosphoglycerate reduction. This inhibition of CO2 fixation appears to be caused by reversal of light induced alkalisation in the stroma by CCCP. 4. Methylamine has a very different effect compared to CCCP. Increasing concentrations of methylamine inhibit CO2 fixation and phosphoglycerate reduction to the same extent. The light induced alkalisation of the stroma appears not to be significantly inhibited by methylamine, but the protons in the thylakoid space are neutralized. The inhibition of CO2 fixation by higher concentrations of methylamine is explained by an inhibition of photophosphorylation. It appears that methylamine does not abolish proton transport. 5. It is shown that intact chloroplasts are able to fix CO2 in the dark, yielding 3-phosphoglycerate. This requires the addition of dihydroxyacetone phosphate as precursor of ribulosemonophosphate and also to supply ATP, and the addition of oxaloacetate for reoxidation of the NADPH in the stroma. 6. Dark CO2 fixation in the presence of dihydroxyacetone phosphate and oxaloacetate has the same pH dependency as CO2 fixation in the light. This demonstrates that CO2 fixation in the dark is not possible, unless the pH in the medium is artificially raised to pH 8.8.  相似文献   

3.
The salts of several weak acids have been used to render the envelope permeable to protons. In order to investigate the role of stromal pH changes in the light regulation of CO2 fixation, formate, octanoate, nitrite, and glyoxylate have been tried as tools to reverse the light-dependent alkalization of the stroma. For this purpose, the decrease of the stromal pH in illuminated spinach chloroplasts, as caused by the addition of these substances or by instantaneous lowering of the pH in the medium, has been compared with the corresponding decrease of CO2 fixation and the change of stromal metabolite levels. It appears from out data that formate and octanoate are suited best to obtain a specific inhibition of CO2 fixation by lowering the stromal pH. The measurement of the corresponding metabolite levels indicates that this inhibition is primarily due to an inhibition of fructose- and sedoheptulose bisphosphatase. It is concluded that these two enzymes are important regulatory steps for the light control of CO2 fixation.  相似文献   

4.
1. The enzymatic steps of the CO2 fixation cycle responsible for the overall inhibition of CO2 fixation caused by the lowering of the Mg2+ concentration in the stroma were investigated. For this the Mg2+ concentration in the stroma was decreased by addition of the ionophore A 23187, and the levels of the intermediates of the CO2 fixation cycle in the stroma of intact chloroplasts were assayed by ion exchange chromatography.2. The addition of the ionophore caused an increase of NADPH, ATP, fructose- and sedoheptulosebisphosphate and a dramatic decrease of phosphoglycerate in the stroma. These changes were reversed by the addition of Mg2+ and again affected by a subsequent addition of Ca2+. Ribulosebisphosphate and pentosemonophosphate levels in the stroma were only a little affected under these different conditions.3. The increase of the NADPH and ATP reflects the decreased utilization of these compounds due to the overall inhibition of CO2 fixation. As phosphoglycerate and triosephosphate appear to be in near equilibrium with NADPH and ATP, the decrease of phosphoglycerate seems to be a consequence of the changes in the nucleotide levels.4. The rapid increase of fructose- and sedoheptulosebisphosphate after the addition of the ionophore A 23187 clearly demonstrates that the overall inhibition of CO2 fixation caused by lowering the stromal Mg2+ is due to the inhibition of the hydrolysis of these sugar bisphosphates. It is concluded that the activities of fructose- and sedoheptulosebisphosphatase can be controlled by light dependent changes of the stromal Mg2+ concentration.  相似文献   

5.
(1) Light-dependent changes of the Mg2+ content of thylakoid membranes were measured at pH 8.0 and compared with earlier measurements at pH 6.6. In a NaCl and KCl medium, the light-dependent decrease in the Mg2+ content of the thylakoid membranes at pH 8.0 is found to be 23 nmol Mg2+ per mg chlorophyll, whereas in a sorbitol medium it is 83 nmol Mg2+ per mg chlorophyll. (2) A light dependent increase in the Mg2+ content of the stroma was detected wjem chloroplasts were subjected to osmotic shock, amounting to 26 nmol/mg chlorophyll. Furthermore, a rapid and reversible light-dependent efflux of Mg2+ has been observed in intact chloroplasts when the divalent cation ionophore A 23 187 was added, indicating a light-dependent transfer of about 60 nmol of Mg2+ per mg chlorophyll from the thylakoid membranes to the stroma. (3) CO2 fixation, but not phosphoglycerate reduction, could be completely inhibited when A 23 187 was added to intact chloroplasts in the absence of external Mg2+. If Mg2+ was then added to the medium, CO2 fixation was restored. Half of the maximal restoration was achieved with about 0.2 mM Mg2+, which is calculated to reflect a Mg2+ concentration in the stroma of 1.2 mM. The further addition of Ca2+ strongly inhibits CO2 fixation. (4) The results suggest that illumination of intact chloroplasts causes an increase in the Mg2+ concentration of 1-3 mM in the stroma. Compared to the total Mg2+ content of chloroplasts, this increase is very low, but it appears to be high enough to have a possible function in the light regulation of CO2 fixation.  相似文献   

6.
EDTA at a concentration of 1 muM produced a threshold effect in the activation of purified rabbit liver fructose-1, 6-diphosphatase [EC 3.1.3.11] in the presence of 5 mM Mg2+ at pH 7.2. Without EDTA, biphasic activation curves were produced by Mg2+. A double-reciprocal plot of the data gave the Km values corresponding to the two linear regions. They were 0.19 and 0.83 mM at pH 7.5, and 0.055 and 0.83 mM at pH 9.1. In the presence of 5muM EDTA a sigmoidal curve was obtained for Mg2+ activation in the range of noninhibitory Mg2+ concentrations at pH 7.2. The apparent Km value for Mg2+ was 0.15 mM, and the Hill coefficient was 2.0. At pH 9.1 cooperativity among the Mg2+ sites disappeared, and the apparent Km value for Mg2+ was 0.055 mM. These Km values at pH 7.2 or 9.1 corresponded to the smaller of the biphasic Km values obtained without EDTA. In the absence of EDTA, no inhibition by Mg2+ was observed in the Mg2+ concentration range below 10 mM. In the presence of EDTA, the enzyme was inhibited markedly by Mg2+ at concentrations above 0.5 mM at pH 7.2, and was more sensitive to inhibition at pH 9.1. The effects of pH on the Km value for Mg2+ activation and on the Mg2+ inhibition contributed to an apparent shift of the pH optimum for activity induced by EDTA. Cooperative interaction among fructose-1, 6-diphosphate sites was observed for the enzyme in the presence of EDTA. The Hill coefficient was approximatley 1.8, and the apparent Km value for the substrate was 0.74 muM. EDTA appears to make liver fructose-1, 6-diphosphatase very sensitive to various effectors. It is suggested that Mg2+ serves as a regulator for the enzyme activity.  相似文献   

7.
1. The pH in the stroma and in the thylakoid space has been measured in a number of chloroplast preparations in the dark and in the light at 20 °C. Illumination causes a decrease of the pH in the thylakoid space by 1.5 and an increase of the pH in the stroma by almost 1 pH unit.2. CO2 fixation is shown to be strongly dependent on the pH in the stroma. The pH optimum was 8.1, with almost zero activity below pH 7.3. Phosphoglycerate reduction, which is a partial reaction of CO2 fixation, shows very little pH dependency.3. Low concentrations of the uncoupler m-chlorocarbonylcyanide phenylhydrazone (CCCP) inhibit CO2 fixation without affecting phosophoglycerate reduction. This inhibition of CO2 fixation appears to be caused by reversal of light induced alkalisation in the stroma by CCCP.4. Methylamine has a very different effect compared to CCCP. Increasing concentrations of methylamine inhibit CO2 fixation and phosphoglycerate reduction to the same extent. The light induced alkalisation of the stroma appears not to be significantly inhibited by methylamine, but the protons in the thylakoid space are neutralized. The inhibition of CO2 fixation by higher concentrations of methylamine is explained by an inhibition of photophosphorylation. It appears that methylamine does not abolish proton transport.5. It is shown that intact chloroplasts are able to fix CO2 in the dark, yielding 3-phosphoglycerate. This requires the addition of dihydroxyacetone phosphate as precursor of ribulosemonophosphate and also to supply ATP, and the addition of oxaloacetate for reoxidation of the NADPH in the stroma.6. Dark CO2 fixation in the presence of dihydroxyacetone phosphate and oxaloacetate has the same pH dependency as CO2 fixation in the light. This demonstrates that CO2 fixation in the dark is not possible, unless the pH in the medium is artificially raised to pH 8.8.7. It is shown that pH changes occurring in the stroma after illumination are sufficient to switch CO2 fixation from zero to maximal activity. This offers a mechanism for light control of CO2 fixation, avoiding wasteful CO2 fixation in the dark.  相似文献   

8.
1. The pH optimum of CO2-dependent O2 evolution by barley (Hordeum vulgare L.) chloroplasts was found to be between 7.8 and 8.2. The addition of 1 mM MgCl2 in the dark inhibited O2 evolution over the entire pH range tested and resulted in a much sharper pH profile centered around pH 8.2. 2. The pH optimum for O2 evolution, in the presence and absence of 1 mM MgCl2, was acid-shifted 0.3--0.4 pH units by 2 mM NH4Cl. The pH optimum of O2 evolution, with and without 1 mM MgCl2, was base-shifted by 2 mM sodium acetate, approx. 0.5 pH units relative to the controls. 3. O2 evolution in the presence of bicarbonate plus 3-phosphoglycerate or ribose-5-phosphate was considerably less sensitive to pH than CO2-dependent O2 evolution in the absence of substrate. With these substrates, both in the presence and absence of 1 mM MgCl2, the pH optimum was broad and was centered around pH 7.8. 4. Inhibition of CO2-dependent O2 evolution by inorganic phosphate and magnesium increased as the pH of the reaction mixture was decreased below the optimum. Decreasing the pH from 8.2 to 7.6, reduced over 3-fold the concentration of inorganic phosphate required to inhibit O2 evolution completely. For magnesium, a similar change in pH reduced the concentration required to inhibit O2 evolution 50% approx. 5-fold. At pH 8.2, magnesium inhibition required inorganic phosphate. Magnesium was not required for inhibition of O2 evolution by inorganic phosphate, but incresaed the relative inhibition observed. 5. Illumination of intact barley chloroplasts increased the activity of NADP-glyceraldehyde-3-P dehydrogenase, phosphoribulokinase and fructose-1,6-diphosphatase. MgCl2 and inorganic phosphate prevented this increase in enzyme activity at concentrations that completely inhibited CO2-dependent O2 evolution. 6. The results obtained suggest that magnesium inhibition of O2 evolution may be caused by enhanced phosphate exchange across the chloroplast envelope.  相似文献   

9.
The sites of action of several compounds, reported to inhibit CO(2) fixation by chloroplast preparations were located by developing assays in lettuce chloroplasts to test their effect on partial reactions of the carbon cycle and on carbonic anhydrase. The results indicated that: d, l-glyceral-dehyde and 5'-AMP inhibit phosphoribulose kinase or isomerase. 3-Phosphoglyceric acid and 6-phosphogluconate inhibit ribulose diphosphate carboxylase. Azide, Mg(2+), and nitrite inhibit the activity of carbonic anhydrase of lettuce chloroplasts and light-dependent CO(2) fixation by intact chloroplasts with similar sensitivities. None of these inhibited CO(2) fixation in ruptured chloroplasts. It is suggested that the inhibition by azide, nitrite, and magnesium ions of CO(2) fixation by intact chloroplasts is due to their inhibition of the activity of carbonic anhydrase.  相似文献   

10.
Reduced ferredoxin: CO2 oxidoreductase (CO2-reductase) from Clostridium pasteurianum catalyzes the reduction of CO2 to formate at the expense of reduced ferredoxin, an isotopic exchange between CO2 and formate in the absence of ferredoxin, and the oxidation of formate to CO2 with oxidized ferredoxin. The three activities were found to be equally affected by monovalent anions known to be ligands to transition metals: The enzyme was reversibly inhibited by azide (Ki = 0.004mM), cyanate (Ki = 0.3 mM), thiocyanate (Ki = 1mM), nitrite (Ki = 0.4mM), nitrate (Ki = 6mM), chlorate (Ki = 3mM), fluoride (Ki = 5mM), and by chloride, bromide, iodide (Ki greater than 5mM). There was no observable effect of pH on the inhibition constants. The enzyme was not inhibited by carbon monoxide. The enzyme was irreversibly inactivated by low concentrations (10muM) of cyanide. The rate of inactivation increased with increasing pH with an inflection point near pH 9.5. Reduced ferredoxin and formate rather than oxidized ferredoxin or CO2 protected the enzyme from inactivation by cyanide. The enzyme was protected by azide and cyanate from inactivation. In the presence of high concentrations of the monovalent anions the rate of inactivation by heat (55 degrees C), by molecular oxygen, and by cyanide was decreased by a factor of more than 100. Half maximal protection was observed at the Ki concentrations of the two reversible inhibitors. The data are interpreted to indicate that a transition metal of weak "a class" character and a disulfide are catalytically significant groups of CO2-reductase from C. pasteurianum.  相似文献   

11.
Sulphite at concentrations from 0.05 to 5.0 mM was supplied to illuminated, detached poplar (Populus deltoides Bart. ex Marsh) leaves via the transpiration stream. The rate of CO2 fixation and partitioning of newly fixed carbon between sucrose and starch were measured and compared with the contents of selected phosphorylated intermediates, the contents of fructose-2,6-bisphosphate (Fru2,6BP) and the activation of sucrose-phosphate synthase (SPS). Supplying leaves with 0.5 mM sulphite led to an increase in the sucrose/starch partitioning ratio without altering the rate of 14CO2 fixation. The increase in sucrose synthesis compared to starch synthesis was accompanied by relatively small changes of 3-phosphoglyceric acid (PGA), fructose-1,6-bisphosphate (Fru1,6BP), hexose phosphates (hexose-)), uridine 5'-diphosphoglucose (UDPGlc), an accumulation of triose phosphates (triose-P), an activation of SPS, and decreased Fru2,6BP contents. Supplying leaves with 1.0 mM sulphite decreased 14CO2 assimilation and increased partitioning of fixed carbon into starch. A selective inhibition of sucrose synthesis was accompanied by an accumulation of triose-P, Fru1,6BP, hexose-P, and a decrease of PGA contents. There was also a large increase of Fru2,6BP contents and a decline in the activation of SPS. It could be argued that sulphite affects the allocation of photosynthetic carbon to sucrose and that sulphite can inhibit photosynthesis via a selective inhibition of sucrose synthesis.  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.
The yields of nitric oxide from 1 mM and 10 mM sodium dithionite in 5 or 150 mM solutions of HEPES buffer (pH 7.4) differed by a factor of 200. Dithionite acted as both a strong reducing agent and an agent responsible for local acidification of the solutions without significant changes in pH. The concentration of nitric oxide was estimated by electron paramagnetic resonance (EPR) by monitoring its incorporation into water-soluble complexes of Fe with N-methyl-D-glucamine dithiocarbamate (MGD), which resulted in the formation of EPR-detectable mononitrosyl complexes of iron. Ten seconds after dithionite addition, the concentration of mononitrosyl iron complexes reached 2 μM, whereas it did not become greater than 0.01 μM in 5 mM HEPES buffer. It has been suggested that this difference results from a longer lifetime of a localized decrease in pH in a weaker buffer solution. This time could be long enough for the protonation of some nitrite molecules. Nitrous acid thus formed decomposed to nitric oxide. A difference in nitric oxide formation from nitrite in weak and strong buffer solutions was also observed in the presence of hemoglobin (0.3 mM) or serum albumin (0.5 mM). However, in the weak buffer the nitric oxide yield was only three-four times greater than in the strong buffer. An increase in the nitric oxide yield from nitrite was observed in solutions containing both proteins. A significant amount of nitric oxide from nitrite was formed in mouse liver preparation subjected to freezing and thawing procedure followed by slurrying in 150 mM HEPES buffer (pH 7.4) and dithionite addition (10 mM). We suggest that the presence of zones with lowered pH values in cells and tissues may be responsible for the predominance of the acidic mechanism of nitric oxide formation from nitrite. The contribution of nitric oxide formation from nitrite catalyzed by heme-containing proteins as nitrite reductases may be minor under these conditions.  相似文献   

14.
The nonglycolytic, anaerobic organism Veillonella parvula M4 has been shown to contain an active pyruvate kinase. The enzyme was purified 126-fold and was shown by disc-gel electrophoresis to contain only two faint contaminating bands. The purified enzyme had a pH optimum of 7.0 in the forward direction and exhibited sigmoidal kinetics at varying concentrations o-f phosphoenol pyruvate (PEP), adenosine 5'-monophosphate (AMP), and Mg-2+ ions with S0.5 values of 1.5, 2.0, and 2.4 mM, respectively. Substrate inhibition was observed above 4 m PEP. Hill plots gave slope values (n) of 4.4 (PEP), 2.8 (adenosine 5'-diphosphate), and 2.0 (Mg-2+), indicating a high degree of cooperativity. The enzyme was inhibited non-competitively by adenosine 5'-triphosphate (Ki = 3.4 mM), and this inhibition was only slightly affected by increasing concentration of Mg-2+ ions to 30 mM. Competitive inhibition was observed with 3-phosphoglycerate, malate, and 2,3-diphosphoglycerate but only at higher inhibitor concentrations. The enzyme was activated by glucose-6-phosphate (P), fructose-6-P, fructose-1,6-diphosphate (P2), dihydroxyacetone-P, and AMP; the Hill coefficients were 2.2, 1.8, 1.5, 2.1, and 2.0, respectively. The presence of each these metabolites caused substrate velocity curves to change from sigmoidal to hyperbolic curves, and each was accompanied by an increase in the maximum activity, e.g., AMP greater than fructose-1,6-P2 greater than dihydroxyacetone-P greater than glucose-6-P greater than fructose-6-P. The activation constants for fructose-1,6-P2, AMP, and glucose-6-P were 0.3, 1.1, and 5.3 mM, respectively. The effect of 5 mM fructose-1,6-P2 was significantly different from the other compounds in that this metabolite was inhibitory between 1.2 and 3 mM PEP. Above this concentration, fructose-1,6-P2 activated the enzyme and abolished substrate inhibition by PEP. The enzyme was not affected by glucose, glyceraldehyde-3-P, 2-phosphoglycerate, lactate, malate, fumerate, succinate, and cyclic AMP. The results suggest that the pyruvate kinase from V. parvula M4 plays a central role in the control of gluconeogenesis in this organism by regulating the concentration of PEP.  相似文献   

15.
Karl-Josef Dietz  Ulrich Heber   《BBA》1984,767(3):432-443
Rates of photosynthesis of spinach leaves were varied by varying light intensity and CO2 concentration. Metabolism of the leaves was then arrested by freezing them in liquid nitrogen. Chloroplasts were isolated by a nonaqueous procedure. In the chloroplast fractions, levels of intermediates of the carbon reduction cycle were determined and considered in relation to the photosynthetic flux situation of the leaves at the time before freezing. During induction of photosynthesis, ribulose 1,5-bisphosphate levels increased in parallel with CO2 fixation. In the steady state, a similar relation between ribulose 1,5-bisphosphate levels and CO2 uptake was observed at light intensities between 0 and 50 W·m−2. A further increase in light intensity increased CO2 fixation rates but not ribulose 1,5-bisphosphate levels. Increasing the CO2 concentration resulted in increased CO2 uptake, whereas ribulose 1,5-bisphosphate levels decreased. Even under CO2 saturation, ribulose 1,5-bisphosphate levels were about 100 nmol/mg chlorophyll corresponding to about 3.5 mM ribulose 1,5-bisphosphate in the chloroplast stroma. This suggests that even under CO2 saturation, ribulose-1,5-bisphosphate carboxylase limits photosynhetic CO2 uptake. Mass action ratios calculated from measured metabolite levels demonstrated that the thermodynamic gradient required for the regeneration of ribulose 1,5-bisphosphate from hexosephosphate and triosephosphate increased considerably as photosynthetic flux increased. Similar calculations revealed that the enzymatic apparatus responsible for the reduction of 3-phosphoglycerate to dihydroxyacetone phosphate is not displaced much from equilibrium even under maximum rates of photosynthesis at saturating CO2. The same is true for aldolase. Fructose-1,6-bisphosphatase also did not limit Calvin cycle turnover. Only at very low light intensities and during the first minutes of the induction period was the ratio of fructose 1,6-bisphosphate to fructose 6-phosphate high. This observation was more readily explained in terms of fructose 1,6-bisphosphate binding to ribulose-1,5-bisphosphate carboxylase than by a rate limitation imposed by insufficient activation of fructose-1,6-bisphosphatase.  相似文献   

16.
R T Proffitt  L Sankaran 《Biochemistry》1976,15(13):2918-2925
Optimal conditions necessary for the reversible inactivation of crystalline rabbit muscle phosphofructokinase by homogeneous rabbit liver fructose-1,6-bisphosphatase have been studied. At higher enzyme levels (to 530 mug/ml of phosphofructokinase) the two proteins were mixed and incubated in a pH 7.5 buffer composed of 50 mM Tris-HC1, 2 mM potassium phosphate, and 0.2 mM dithiothreitol. Aliquots were removed at various times and assayed for enzyme activity. A time dependent inactivation of phosphofructokinase caused by 1-2.3 times its weight of fructose-1,6-bisphosphatase was observed at 30, 23, and 0 degree C. This inactivation did not require the presence of adenosine 5'-triphosphate or Mg2+ in the incubation mixture, but an adenosine 5'-triphosphate concentration of 2.7 mM or greater was required in the assay to keep phosphofructokinase in an inactive form. A mixture of activators (inorganic phosphate, (NH4)2SO4, and adenosine 5'-monophosphate), when added to the assay cuvette, restored nearly all of the expected enzyme activity. Incubations with other proteins, including aldolase, at concentrations equal to or greater than the effective quantity of fructose-1,6-bisphosphatase had no inhibitory effect on phosphofructokinase activity. Removal of tightly bound fructose 1,6-bisphosphate from phosphofructokinase could not explain this inactivation, since several analyses of crystalline phosphofructokinase averaged less than 0.1 mol of fructose 1,6-bisphosphate/320 000 g of enzyme. Furthermore, the inactivation occurred in the absence of Mg2+ where the complete lack of fructose-1-6-bisphosphatase activity was confirmed directly. At lower phosphofructokinase concentrations (0.2-2 mug/ml) the inactivation was studied directly in the assay cuvette. Higher ratios of fructose-1,6-bisphosphatase to phosphofructokinase were necessary in these cases, but oleate and 3-phosphoglycerate acted synergistically with lower amounts of fructose-1,6-bisphosphatase to cause inactivation. The inactivation did not occur when high concentrations of fructose 6-phosphate were present in the assay, or when the level of adenosine 5'-triphosphate was decreased. However, the inactivation was found at pH 8, where the effects of allosteric regulators on phosphofructokinase are greatly reduced. Experiments with rat liver phosphofructokinase showed that this enzyme was also subject to inhibition by rabbit liver fructose 1,6-bisphosphatase under conditions similar to those used in the muscle enzyme studies. Attempts to demonstrate direct interaction between phosphofructokinase and fructose-1,6-bisphosphate by physical methods were unsuccessful. Nevertheless, our results suggest that, under conditions which approximate the physiological state, the presence of fructose-1,6bisphosphatase can cause phosphofructokinase to assume an inactive conformation. This interaction may have a significant role in vivo in controlling the interrelationship between glycolysis and gluconeogenesis.  相似文献   

17.
(1) Light-dependent changes of the Mg2+ content of thylakoid membranes were measured at pH 8.0 and compared with earlier measurements at pH 6.6. In a NaCl and KCl medium, the light-dependent decrease in the Mg2+ content of the thylakoid membranes at pH 8.0 is found to be 23 nmol Mg2+ per mg chlorophyll, whereas in a sorbitol medium it is 83 nmol Mg2+ per mg chlorophyll.

(2) A light dependent increase in the Mg2+ content of the stroma was detected when chloroplasts were subjected to osmotic shock, amounting to 26 nmol/mg chlorophyll. Furthermore, a rapid and reversible light-dependent efflux of Mg2+ has been observed in intact chloroplasts when the divalent cation ionophore A 23 187 was added, indicating a light-dependent transfer of about 60 nmol of Mg2+ per mg chlorophyll from the thylakoid membranes to the stroma.

(3) CO2 fixation, but not phosphoglycerate reduction, could be completely inhibited when A 23 187 was added to intact chloroplasts in the absence of external Mg2+. If Mg2+ was then added to the medium, CO2 fixation was restored. Half of the maximal restoration was achieved with about 0.2 mM Mg2+, which is calculated to reflect a Mg2+ concentration in the stroma of 1.2 mM. The further addition of Ca2+ strongly inhibits CO2 fixation.

(4) The results suggest that illumination of intact chloroplasts causes an increase in the Mg2+ concentration of 1–3 mM in the stroma. Compared to the total Mg2+ content of chloroplasts, this increase is very low, but it appears to be high enough to have a possible function in the light regulation of CO2 fixation.  相似文献   


18.
Nitrite is the highly toxic end product of ammonia oxidation that accumulates in the absence of a nitrite-consuming process and is inhibitory to nitrifying and other bacteria. The effects of nitrite on ammonia oxidation rates and regulation of a common gene set were compared in three ammonia-oxidizing bacteria (AOB) to determine whether responses to this toxic metabolite were uniform. Mid-exponential-phase cells of Nitrosomonas europaea ATCC 19718, Nitrosospira multiformis ATCC 25196, and Nitrosomonas eutropha C-91 were incubated for 6 h in mineral medium supplemented with 0, 10, or 20 mM NaNO(2) . The rates of ammonia oxidation (nitrite production) decreased significantly only in NaNO(2) -supplemented incubations of N. eutropha; no significant effect on the rates was observed for N. europaea or N. multiformis. The levels of norB (nitric oxide reductases), cytL (cytochrome P460), and cytS (cytochrome c'-β) mRNA were unaffected by nitrite in all strains. The levels of nirK (nitrite reductase) mRNA increased only in N. europaea in response to nitrite (10 and 20 mM). Nitrite (20 mM) significantly reduced the mRNA levels of amoA (ammonia monooxygenase) in N. multiformis and norS (nitric oxide reductase) in the two Nitrosomonas spp. Differences in response to nitrite indicated nonuniform adaptive and regulatory strategies of AOB, even between closely related species.  相似文献   

19.
The unresolved autotrophic CO2 fixation pathways in the sulfur-reducing Archaebacterium Thermoproteus neutrophilus and in the phototrophic Eubacterium Chloroflexus aurantiacus have been investigated. Autotrophically growing cultures were labelled with [1,4-13C1]succinate, and the 13C pattern in cell constituents was determined by 1H- and 13C-NMR spectroscopy of purified amino acids and other cell constituents. In both organisms succinate contributed to less than 10% of cell carbon, the major part of carbon originated from CO2. All cell constituents became 13C-labelled, but different patterns were observed in the two organisms. This proves that two different cyclic CO2 fixation pathways are operating in autotrophic carbon assimilation in both of which succinate is an intermediate. The 13C-labelling pattern in T. neutrophilus is consistent with the operation of a reductive citric acid cycle and rules out any other known autotrophic CO2 fixation pathway. Surprisingly, the proffered [1,4-13C1]succinate was partially converted to double-labelled [3,4-13C2]glutamate, but not to double-labelled aspartate. These findings suggest that the conversion of citrate to 2-oxoglutarate is readily reversible under the growth conditions used, and a reversible citrate cleavage reaction is proposed. The 13C-labelling pattern in C. aurantiacus disagrees with any of the established CO2 fixation pathways; it therefore demands a novel autotrophic CO2 fixation cycle in which 3-hydroxypropionate and succinate are likely intermediates. The bacterium excreted substantial amounts of 3-hydroxypropionate (5 mM) and succinate (0.5 mM) at the end of autotrophic growth. Autotrophically grown Chloroflexus cells contained acetyl-CoA carboxylase and propionyl-CoA carboxylase activity. These enzymes are proposed to be the main CO2-fixing enzymes resulting in malonyl-CoA and methylmalonyl-CoA formation; from these carboxylation products 3-hydroxypropionate and succinate, respectively, can be formed.  相似文献   

20.
Pyruvate Kinase of Streptococcus lactis   总被引:18,自引:14,他引:4       下载免费PDF全文
The kinetic properties of pyruvate kinase (ATP:pyruvate-phosphotransferase, EC 2.7.1.40) from Streptococcus lactis have been investigated. Positive homotropic kinetics were observed with phosphoenolpyruvate and adenosine 5′-diphosphate, resulting in a sigmoid relationship between reaction velocity and substrate concentrations. This relationship was abolished with an excess of the heterotropic effector fructose-1,6-diphosphate, giving a typical Michaelis-Menten relationship. Increasing the concentration of fructose-1,6-diphosphate increased the apparent Vmax values and decreased the Km values for both substrates. Catalysis by pyruvate kinase proceeded optimally at pH 6.9 to 7.5 and was markedly inhibited by inorganic phosphate and sulfate ions. Under certain conditions adenosine 5′-triphosphate also caused inhibition. The Km values for phosphoenolpyruvate and adenosine 5′-diphosphate in the presence of 2 mM fructose-1,6-diphosphate were 0.17 mM and 1 mM, respectively. The concentration of fructose-1,6-diphosphate giving one-half maximal velocity with 2 mM phosphoenolpyruvate and 5 mM adenosine 5′-diphosphate was 0.07 mM. The intracellular concentrations of these metabolites (0.8 mM phosphoenolpyruvate, 2.4 mM adenosine 5′-diphosphate, and 18 mM fructose-1,6-diphosphate) suggest that the pyruvate kinase in S. lactis approaches maximal activity in exponentially growing cells. The role of pyruvate kinase in the regulation of the glycolytic pathway in lactic streptococci is discussed.  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