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.
The effectr of phosphate starvation and subsequent uptake on distribution and concentration of phosphate metabolic intermediates and metals were studied in Heterosigma akashiwo (Hada) Hada by 31P-NMR spectroscopy, neutron activation analysis and ESR spectroscopy. Excess orthophosphate (4.5 μM Pi, as NaH2PO4) added to a medium with P-depleted H. akashiwo cells was rapidly taken up resulting in an increase in P cell quota (qp)from 68.2 to 99.6 fmol. cell-1in 2 h and to 156.3 fmol. cell-1in 6 h. After three days, qp approached about 190 fmol. cell?1. Polyphosphate (PPi) rapidly increased from 0 to 11.4 fmol· cell?1in 2 h and to 24.7 fmol·cell?1in 6 h. Diel variation of cell quota indicated that cellular Pi increase was synchronized with cellular PPi decrease and vice versa. The average chain length of PPi increased from ca. 0 to ca. 10.2 phosphate residues in 2 h after addition of Pi and one day later, from ca. 9.8 to ca. 12.5. The cell quota of Mn (qMn), and to a lesser extent Co, increased rapidly from 4.87 fg. cell?1in the P- starved condition to 50.48 fg·cell?12 h afer addition of Pi but decreased to 8.63 fg. Cell?1by 6 h. Concentrations of Zn, As, Hf, Cu and sometimes Al, Mg, K, and Ca changed in a manner opposite to that of Mn and Co. The excretion of these cations, which was synchronized with the uptake of Mn and Co, may be important for a charge balancing in the cells. The ESR spectra showed that the high cellular Mn observed at 2 h after P addition was Mn2+which was taken up by the cells rather than adsorbed on the cell surface. These data combined with PPi data suggested that the behavior of qMn is synchronized with the behavior of average chain length of PPi.  相似文献   

2.
The effects of starvation and subsequent addition of phosphate-containing medium on the phosphate metabolic intermediates were studied by 31P-NMR spectroscope of perchloric acid extracts and intact cells of Heterosigma akashiwo (Hada) Hada. When orthophosphate in the medium was completely depleted the medium was enriched with orthophosphate (4.5 μM). In the phosphate starved condition, the P cell quota was 76 fmol·cell−1 and the major components of phosphate intermediates were phosphodiester, sugar phosphate and orthophosphate (Pi). After addition of Pi, rapid uptake of Pi was observed and the P cell quota increased to 108 fmol·cell−1 in 2 h, 134 fmol·cell−1 in 5 h and 222 fmol·cell−1 in 1 day after addition of phosphate. The 31P-NMR spectrum indicated that a major portion of P was stored as polyphosphate, in which the average chain length of polyphosphate increased from 10 to 20 phosphate residues in one day after addition of Pi.  相似文献   

3.
The effects of starvation and subsequent addition of phosphate-containing medium on the phosphate metabolic intermediates were studied by 31P-NMR spectroscopy of perchloric acid extracts and intact cells of Heterosigma akashiwo (Hada) hada. When orthophosphate in the medium was completely depleted the medium was enriched with orthophosphate (4.5 μM). In the phosphate starved condition, the P cell quota was 76 fmol-cell−1 and the major components of phosphate intermediates were phosphodiester, sugar phosphate and orthophosphate (Pi) After addition of Pi' rapid uptake of Pi was observed and the P cell quota increased to 108 fmol. cell−1 in 2 h, 134 fmol. cell−1 in 5 h and 222 fmol. cell−1 in 1 day after addition of phosphate. The 31P-NMR spectrum indicated that a major portion of P was stored as polyphosphate, in which the average chain length of polyphosphate increased from 10 to 20 phosphate residues in one day after addition of Pi-  相似文献   

4.
Two planktonic algal species, Staurastrum chaetoceras (Schr.) G. M. Smith and Cosmarium abbreviatum Rac. var. planctonicum W. et G. S. West, from trophically different alkaline lakes, were compared in their response to a single saturating addition of phosphate (P) in a P-limited growth situation. Storage abilities were determined using the luxury coefficient R = Qmax/Q0. Maximum cellular P quotas differed, depending on whether cells were harvested during exponential growth at μmax (Qmax, R being 26.7 and 9.1 for C. abbreviatum and S. chaetoceras, respectively) or harvested after a saturating pulse at P-limited growth conditions (Q′max, R being 53.5 and 20.2 for C. abbreviatum and S. chaetoceras, respectively). At stringent P-limited conditions, maximum initial uptake rates were higher in S. chaetoceras than in C. abbreviatum (0.094 and 0.073 pmol P·cell?1·h?1, respectively), but long-term (net) uptake rates (over ~20 min) were higher in C. abbreviatum than in S. chaetoceras (0.048 and 0.019 pmol P·cell?1·h?1, respectively). Before growth resumed after the onset of a large P addition (150 μmol·L?1), a lag phase was observed for both species. This period lasted 2–3 days for S. chaetoceras and 3–4 days for C. abbreviatum, corresponding with the time to reach Qmax. Subsequent growth rates (over ~10 days) were 0.010 h?1 and 0.006 h?1 for S. chaetoceras and C. abbreviatum, respectively, being only 20%–30% of maximum growth rates. In conclusion, S. chaetoceras, with a relatively high initial P-uptake rate, short lag phase, and high initial growth rate, is well adapted to a P pulse of short duration. Conversely, C. abbreviatum, with a high long-term uptake rate and high storage capacity, appears competitively superior when exposed to an infrequent but lasting pulse. These characteristics provide information about possible strategies of algal species to profit from temporarily high P concentrations.  相似文献   

5.
The desmid Staurastrum luetkemuellerii Donat et Ruttner and the cyanobacterium Microcystis aeruginosa Kütz. showed pronounced differences in chemical composition and ability to maintain P fluxes. The cellular P:C ratio (Qp) and the surplus P:C ratio (Qsp) were higher in M. aeruginosa, indicating a lower yield of biomass C per unit of P. The subsistence quota (Qp) was 1.85 μg P·mg C?1in S. luetkemuellerii and 6.09 μg P·mg C?1in M. aeruginosa, whereas the respective Qp of P saturnted organisms (Qs) were 43 and 63 μg P·mg C?1. These stores could support four divisions in S. luetkemuellerii and three divisions in M. aeruginosa, which suggests that the former exhibited highest storage capacity (Qs/Q0). M. aeruginosa showed a tenfold higher activity of alkaline phosphatase than S. luetkemuellerii when P starved. The optimum N:P ratio (by weight) was 5 in S. luetkemuellerii and 7 in M. aeruginosa. The initial uptake of Pi pulses in the organisms was not inhibited by rapid (<1 h) internal feedback mechanisms and the short term uptake rote could be expressed solely as a function of ambient Pi. The maximum cellular C-based uptake rate (Vm) in P starved M. aeruginosa was up to 50 times higher than that of S. luetkemuellerii. It decreased with increasing growth rate (P status) in the former species and remained fairly constant in the latter. The corresponding cellular P-based value (Um= Vm/Qp) decreased with growth rate in both species and was about 10 times higher in P started M. aeruginosa than in S. luetkemuellerii. The average half saturation constant for uptake (Km) was equal for both species (22 μg P·L?1) and varied with the P status. S. luetkemuellerii exhibited shifts in the uptake rate of Pi that were characterized by increased affinity (Um/Km) at low Pi, concentrations (<4 μg P·L?1) compared to that at higher concentrations. The species thus was well adapted to uptake at low ambient Pi, but M. aeruginosa was superior in Pi uptake under steady state and transient conditions when the growth rate was lower than 0.75 d?1. Moreover, M. aeruginosa was favored by pulsed addition of Pi. M. aeruginosa relpased Pi at a higher rate than S. luetkemuellerii. Leakage of Pi from the cells caused C-shaped μ vs. Pi curves. Therefore, no unique Ks for growth could be estimated. The maximum growth rate (μm) (23° C) was 0.94 d?1for S. luetkemuellerii and 0.81 d?1for M. aeruginosa. The steady state concentration of Pi (P*) was lower in M. aeruginosa than in S. luetkemuellerii at medium growth rates. The concentration of Pi at which the uptake and release of Pi was equal (Pc was, however, lower in S. luetkemuellerii.  相似文献   

6.
Diel vertical migration by Heterosigma akashiwo (Hada) Hada (Raphidophyceae) was monitored in a 1.5 in tall microcosm. Vertical stratification, with low salinity and low orthophosphate (Pi) concentration in the upper layer and high salinity and high Pi concentration in the lower layer, was simulated in the tank, analogous to summer stratification in the Seto Inland Sea. The phosphate metabolism of H. akashiwo during this vertical migration was studied using 31P-NMR spectroscopy. At night this species migrated to the lower phosphate-rich layer and took up inorganic phosphate (Pi) which then was accumulated as polyphosphate (PPi) by an increase in the chain length of PPi During the daytime this species migrated to the phosphate-depleted surface water and utilized the accumulated PPi for photophosphorylation by decreasing the chain length of PPi During the first night after the phosphorus was introduced to the previously impoverished waters, the cells took up inorganic phosphate, accumulating the new phosphorus nutrient internally as Pi But the cells did not convert Pi to PPi presumably due to their lack of ATP. After the second day of the experiment, conversion of Pi to PPi at night was much more rapid than on the first day, presumably due to increased ATP availability. Then the cycle continued, with uptake of Pi and conversion to PPi at night at the bottom and its utilization during the day at the surface. These data suggest that the role of PPi in the metabolism of this species appears to be as a phosphate pool which regulates the level of Pi and ATP in the cell. Diel vertical migration allows this red tide species to shuttle between the phosphate-rich lower layer and the photic upper layer in stratified waters. 31P-NMR is shown to be a valuable tool in studying the phosphorus metabolism in migrating organisms.  相似文献   

7.
In studying conditions for obtaining photosynthetically functional chloroplasts from mesophyll protoplasts of sunflower and wheat, a strong requirement for chelation was found. The concentration of chelator, either EDTA or pyrophosphate (PPi), required for maximum activation depended on the pH, the concentration of orthophosphate (Pi) in the assay, and the chelator used. Studies with EDTA indicate that including the chelator in the isolation, resuspension, and assay media, in the absence of divalent cations, was most effective. Increased concentration of EDTA from 1 to 10 mm broadened the pH response curve for photosynthesis, inasmuch as a higher concentration of chelator was required for activation of photosynthesis at lower pH.Either EDTA, PPi, or citrate could activate photosynthesis of sunflower chloroplasts isolated and assayed at pH 8.4. At pH 7.6, PPi and EDTA were equally effective at low Pi concentrations but PPi was particularly effective in shortening the induction period at high concentrations of Pi (2.5 mm) in the assay medium. Including 1 mm 3-phosphoglycerate in the assay medium with or without Pi could not replace the need for chelation. However, 3-phosphoglycerate + EDTA in the assay medium with 0.5 mm Pi, pH 7.6, gave a short induction period and rates of photosynthesis similar to those with 10 mm PPi. The results suggest that PPi can have a dual effect at the lower pH through chelation and inhibition of the phosphate transporter.Photosynthesis by sunflower chloroplasts isolated and assayed at pH 8.4 with 0.2 mm EDTA (+ 0.5 mm Pi in the assays) was severely inhibited by 2 mM CaCl2, MgCl2, or MnCl2. Wheat chloroplasts isolated and assayed at pH 8.4 without chelation, and assayed with 0.2 mm Pi, had low rates of photosynthesis (25 μmol O2 evolved mg?1 chlorophyll h?1) which were strongly inhibited by 2 to 4 mm MgCl2, MnCl2, or CaCl2. With inclusion of EDTA and Pi at optimum levels, isolated chloroplasts of sunflower and wheat have high rates of photosynthesis and PPi or divalent cations are not of benefit.  相似文献   

8.
The inorganic phosphorus (Pi) uptake kinetics of Spirogyra fluviatilis Hilse were examined as a function of phosphorus cell quota (QP) and flow velocity in a laboratory stream apparatus. Short-term uptake and the acclimation of the uptake mechanism to flow were measured by the disappearance of Pi pulses in a recirculating flow cell. Short-term Pi uptake was biphasic. When the alga was P-deficient, Phase 1 and 2 half-saturation constants and maximum uptake rates were 11.0 and 47.2 μg P·L?1 and 473 and 803 μg P·g dry wt?1 h?1, respectively. Flowing water altered short-term uptake when the alga was P-deficient, but not when it was P-replete. When QP was less than 0.21%, increases in flow velocity from 3 to 15 cm·s?1 enhanced uptake with maximum uptake for any Pi pulse at 12 and 15 cm·s?1. At 22 and 30 cm·s?1, uptake was reduced by 12% or more relative to the maxima. If, however, the alga was cultivated at 22 and 30 cm·s?1 and short-term Pi uptake was measured at 12 cm·s?1, uptake was on average 33% greater than when the alga was cultivated at the latter velocity. Apparently, the alga could adjust short-term uptake to compensate for the suboptimal conditions of the faster velocities. Long-term Pi uptake and net phosphorus efflux were estimated by a non-steady state application of the Droop equation. Long-term uptake of very low Pi concentrations was not reduced by fast flowing water. Instead, uptake increased proportionately with flow velocity. Maximum phosphorus efflux from S. fluviatilis was 3% of cellular P per hour and occurred when QP was greater than 0.2%. At lower QP, the hourly efflux rate was typically less than 1%. Flowing water did not greatly enhance efflux, although when Pi was undetectable, efflux did tend to increase slightly with velocity. The data show that the effects of flowing water on Pi uptake were varied and not always beneficial. If the effects of flowing water on nutrient acquisition by other lotic algae are similarly varied and complex, flow may be an important determinant of nutrient partitioning among benthic algae in streams.  相似文献   

9.
Emiliania huxleyi (strain L) expressed an exceptional P assimilation capability. Under P limitation, the minimum cell P content was 2.6 fmol P·cell?1, and cell N remained constant at all growth rates at 100 fmol N·cell?1. Both, calcification of cells and the induction of the phosphate uptake system were inversely correlated with growth rate. The highest (cellular P based) maximum phosphate uptake rate (VmaxP) was 1400 times (i.e. 8.9 h?1) higher than the actual uptake rate. The affinity of the P‐uptake system (dV/dS) was 19.8 L·μmol?1·h?1 at μ = 0.14 d?1. This is the highest value ever reported for a phytoplankton species. Vmax and dV/dS for phosphate uptake were 48% and 15% lower in the dark than in the light at the lowest growth rates. The half‐saturation constant for growth was 1.1 nM. The coefficient for luxury phosphate uptake (Qmaxt/Qmin) was 31. Under P limitation, E. huxleyi expressed two different types of alkaline phosphatase (APase) enzyme kinetics. One type was synthesized constitutively and possessed a Vmax and half‐saturation constant of 43 fmol MFP·cell?1·h?1 and 1.9 μM, respectively. The other, inducible type of APase expressed its highest activity at the lowest growth rates, with a Vmax and half‐saturation constant of 190 fmol MFP·cell?1·h?1 and 12.2 μM, respectively. Both APase systems were located in a lipid membrane close to the cell wall. Under N‐limiting growth conditions, the minimum N quotum was 43 fmol N·cell?1. The highest value for the cell N‐specific maximum nitrate uptake rate (VmaxN) was 0.075 h?1; for the affinity of nitrate uptake, 0.37 L·μmol?1·h?1. The uptake rate of nitrate in the dark was 70% lower than in the light. N‐limited cells were smaller than P‐limited cells and contained 50% less organic and inorganic carbon. In comparison with other algae, E. huxleyi is a poor competitor for nitrate under N limitation. As a consequence of its high affinity for inorganic phosphate, and the presence of two different types of APase in terms of kinetics, E. huxleyi is expected to perform well in P‐controlled ecosystems.  相似文献   

10.
Summary Genetic studies suggest that the so-called phosphorus-family of enzymes inN. crassa are controlled by a complex system of regulatory genes which are responsive to the level of phosphorus in the growth medium. The intracellular metabolite(s) that interact with this system to signal changes in the external phosphorus concentration has not been identified. In this study the pools of acid-soluble, phosphorus-containing, compounds are measured in wild-type and phosphorus-family enzyme regulatory mutant strains ofN. crassa before and during phosphorus starvation.Prolonged phosphorus starvation of wild-typeN. crassa failed to alter significantly the pre-starvation level of intracellular orthophosphate, suggesting that intracellular Pi would be a poor effector signal for the control of the phosphorus family enzymes. However, inorganic pyrophosphate (PPi) decreased 15-fold, and tri- and tetrapolyphosphate (PPPi and PPPPi) increased 3- to 5-fold within 15 minutes after transfer of the wild-type strain to phosphorus-free medium. Phosphate starvation of seven different regulatory gene mutant strains resulted in a rapid decrease in the PPi pool similar to that which occurred in the wild-type. However, only two of these seven strains showed increased PPPi and PPPPi pools following phosphate starvation. Additional experiments demonstrated that PPi pools, but not PPPi and PPPPi pools, were unaffected by several starvation regimens other than phosphorus starvation. Metabolic studies employing H3 32PO4 showed that the pool of PPi was labeled to steady-state levels after two minutes of continuous labeling of a phosphate-sufficient culture. Furthermore, long-term steady-state labeling showed that the intracellular PPi pool was directly responsive to the decrease in the extracellular Pi concentration of the medium resulting from cell growth. Growth on phosphoethanolamine, a phosphorus source that allows a modest degree of derepression even in growing cells, resulted in lower levels of PPi than were seen in phosphate-grown cells. These observations suggest that PPi may be involved in the mechanism responsible for the control of phosphorus-family enzyme regulatory gene product activity.  相似文献   

11.
Gerri Levine  J.A. Bassham 《BBA》1974,333(1):136-140
Inhibition of photosynthesis in isolated spinach chloroplasts by Pi is decreased by the presence of PPi and increased with increasing Mg2+ concentration. Previously reported regulation of this photosynthesis by protein factors from spinach leaves appears to be due mostly to pyrophosphate phosphohydrolase (EC 3.6.1.1) activity which converts PPi to Pi and to the effects of PPi and Mg2+ on this pyrophosphatase activity.  相似文献   

12.
Monodentate Co(NH3)5PPi was determined not to be a substrate for yeast inorganic pyrophosphatase while P1,P2-bidentate Co(NH3)4PPi was turned over by the enzyme at a rate of 7.5 min?1. A kinetic analysis of the substrate activities of the P1,P2-bidentate complexes, Co(en)2PPi, Cr(NH3)4PPi, Cr(H2O)(NH3)3PPi, Cr(H2O)2(NH3)2PPi, and Cr(H2O)4PPi was carried out in order to access the potential role of the metal-water ligands in productive binding. While substitution of the H2O ligands with NH3 ligands had a minimal affect on the Km for Mg2+, the binding affinity of the complexes decreased with an increasing NH3H2O ligand ratio as did the turnover number of the corresponding central complexes. The Co(en)2PPi complex was hydrolyzed at a rate approximately 0.6% of that for the Co(NH3)4PPi complex. The substrate activities of β,γ-bidentate Co(NH3)4PPPi and α,β,γ-tridentate Co(NH3)3PPP with pyrophosphatase were also tested. While both complexes were shown to bind tightly to the Mg2+-activated enzyme neither was hydrolyzed. On the other hand, in the presence of the Zn2+-activated enzyme the tridentate complex was turned over at a rate of 0.17 min?1 while the bidentate complex remained inert to hydrolysis.  相似文献   

13.
A microprocedure for the colorimetric determination of inorganic pyrophosphate (PPi) in the presence or absence of orthophosphate (Pi) has been developed. PPi is estimated quantitatively as the amount of chromophore formed with molybdate reagent, 1-amino-2-naphthol-4-sulfonic acid in bisulfite and thiol reagent (monothioglycerol or 2-mercaptoethanol). The latter is obligatory for color formation. Pi is estimated without thiol reagent. The two chromophores differ in absorption spectra, the greatest difference being at 580 nm. For both, color develops fully by 10 min and is stable up to 1 hr. Just less than 0.4 μm PPi can be detemined. The extinction coefficients are 2.70 × 104 and 8.76 × 103 for PPi and Pi, respectively, both with thiol reagent present, and 2.77 × 103 for Pi with no thiol reagent.A ten-fold excess of Pi does not interfere with the determination of PPi and in fact can be estimated in the same mixture. A 15-fold excess, however, diminishes the accuracy of PPi estimations. Trichloroacetic acid and sodium fluoride inhibi color formation, but this inhibition is overcome by the addition of sodium acetate buffer, pH 4.0. Nucleoside triphosphates and adenosine 3′:5′-cyclic monophosphate are stable in the reaction mixture.The method was tested in assays of Escherichia coli DNA-dependent RNA polymerase (nucleoside triphosphate: RNA nucleotidyltransferase, EC 2.7.7.6). Progress curves measured by either the rate of PPi formation or the rate of synthesis of labeled RNA were very similar. Product PPi formed by as little as 0.6 unit of RNA polymerase in a 225-μl incubation medium could be measured.An automated version of the method was devised which allows accurate determination of PPi down to 1 μm (without range expander attachment) at a sampling rate of 20–40 tubes/hr.  相似文献   

14.
The effects of NH4+ assimilation on dark carbon fixation and β-1,3-glucan metabolism in the N-limited marine diatom Skeletonema costatum (Grev.) Cleve (Bacillariophyceae) were investigated by chemical analysis of cell components and incorporation of 14C-bicarbonate. The diatom was grown in pH-regulated batch cultures with a 14:10 h LD cycle until N depletion. The cells were then incubated in the dark with 14C-bicarbonate, but without a source of N for 2 h, then in the dark with 63 μmol·L−1 NH4+ for 3 h. Without N, the cellular concentration of free amino acids was almost constant (∼4.5 fmol·cell−1). Added NH4+ was assimilated at a rate of 12 fmol·cell−1·h−1, and the cellular amino acid pool increased rapidly (doubled in <1 h, tripled in <3 h). The glutamine level increased steeply (45× within 3 h), and the Gln/ Glu ratio increased from 0.1 to 2.4 within 3 h. The rate of dark C fixation during N depletion was only 1.0 fmol·cell−1·h−1. The addition of NH4+ strongly stimulated dark C fixation, leading to an assimilation rate of 4.0 fmol·cell−1·h−1, corresponding to a molar C/N uptake ratio of 0.33. Biochemical fractionation of organic 14C showed no significant 14C fixation into amino acids during N depletion, but during the first 1–2 h of NH4+ assimilation, amino acids were rapidly radiolabeled, accounting for virtually all net 14C fixation. These results indicate that anaplerotic β-carboxylation is activated during NH4+ assimilation to provide C4 intermediates for amino acid biosynthesis. The level of cellular β-1,3-d-glucan was constant (16.5 pg·cell−1) during N depletion, but NH4+ assimilation activated a mobilization of 28% of the reserve glucan within 3 h. The results indicate that β-1,3-glucan in diatoms is the ultimate substrate for β-carboxylation, providing precursors for amino acid biosynthesis in addition to energy from respiration.  相似文献   

15.
Paralytic shellfish toxins, pigment composition, and large subunit (LSU) rDNA sequence were analyzed for a clonal culture of Alexandrium minutum Halim isolated in 2000 from the coastal Fleet Lagoon, Dorset, United Kingdom. The HPLC pigment analysis revealed the presence of chl a, peridinin, and diadinoxanthin as major pigments and chl c1+c2 and c3, diatoxanthin, and β‐carotene as minor components. The toxins responsible for paralytic shellfish poisoning were analyzed by HPLC with postcolumn derivatization and fluorescence detection. The paralytic shellfish poisoning toxin profile of the Fleet Lagoon strain of A. minutum in exponential growth phase was dominated by gonyautoxin‐3 up to 54%, whereas gonyautoxin‐2 made up 10% and saxitoxin (STX) 36%. The average toxicity of the culture was 3.8 pg STX Eq·cell?1, and total toxin content varied from 5.6 fmol·cell?1 on day 1 to a maximum of 16.8 fmol·cell?1 during the early stationary phase. Sequence analysis of the LSU rDNA revealed the strain to be closely related to several European strains of A. minutum and one isolated from Australian waters, although most of these do not produce STX. The shallow Fleet Lagoon may provide a favorable environment for A. minutum to bloom, and the presence of highly potent saxitoxins in this strain indicates potential for future shellfish contamination.  相似文献   

16.
A generally applicable, inexpensive, and sensitive method for the determination of inorganic pyrophosphate (PPi) was developed. PPi was quantitatively separable from solution even in nanomolar concentrations by filtration through a membrane filter in the presence of CaCl2 and KF. The separated PPi was dissolved by immersing the filter in 0.5 n H2SO4. Inorganic phosphate (Pi) was removed by precipitating it as a phosphomolybdate-triethylamine complex and the PPi was measured as a green pyrophosphomolybdate in the presence of 2-mercaptoethanol. Nucleotides and phosphate esters do not react. PPi can be accurately assayed even when there is a 104-fold excess of Pi. Trimetaphosphate, tripolyphosphate, and tetrapolyphosphate also give this green color, but the rate of the color formation is 50 times slower than that with PPi. Thus this interference of the polyphosphates can be eliminated or the polyphosphates can be assayed simultaneously with the PPi in the same sample.  相似文献   

17.
Although the capacity of isolated β-subunits of the ATP synthase/ATPase to perform catalysis has been extensively studied, the results have not conclusively shown that the subunits are catalytically active. Since soluble F1 of mitochondrial H+-ATPase can bind inorganic pyrophosphate (PPi) and synthesize PPi from medium phosphate, we examined if purified His-tagged β-subunits from Thermophilic bacillus PS3 can hydrolyze PPi. The difference spectra in the near UV CD of β-subunits with and without PPi show that PPi binds to the subunits. Other studies show that β-subunits hydrolyze [32P] PPi through a Mg2+-dependent process with an optimal pH of 8.3. Free Mg2+ is required for maximal hydrolytic rates. The Km for PPi is 75 μM and the Vmax is 800 pmol/min/mg. ATP is a weak inhibitor of the reaction, it diminishes the Vmax and increases the Km for PPi. Thus, isolated β-subunits are catalytically competent with PPi as substrate; apparently, the assembly of β-subunits into the ATPase complex changes substrate specificity, and leads to an increase in catalytic rates.  相似文献   

18.
The effects of flowing water on net photosynthesis, dark respiration, specific growth rate, and optimum N:P ratios by Spirogyra fluviatilis Hilse were assessed. The alga was cultivated under nitrogen or phosphorus limitation in laboratory streams at three flow velocities: 3, 12, and 30 cm·s?1. The Droop equation adequately described respiration and photosynthesis (PSnet) as a function of N or P cell quota (QN or Qp). The data show that for N- or P-limited Spirogyra fluviatilis, flowing water is physiologically costly. Generally, flowing water had little effect on respiration rates; however, the proportion of gross photosynthesis devoted to dark respiration did increase with flow velocity. For photosynthesis, the minimum N and P cell quotas increased with velocity, and the theoretical PSnet maxima for N and P both appeared greatest at 12 cm·s?1. The Droop models showed that for any given QN or Qp, PSnet, was reduced by the 30-cm·s?1 treatment. Consistent with this finding, independent estimates of specific growth rates for P-limited S. fluviatilis in the laboratory streams were inversely related to flow velocity when ambient PO4?3 was undetectable. However, growth was not diminished at the fastest velocity when PO4?3 was available for uptake. Thus, the increase in cellular phosphorus demand can be offset by flow-enhanced P uptake when conditions permit; otherwise, growth will be impaired. The optimum N:P ratios for S. fluviatilis at 3, 12, and 30 cm·s?1 were 50, 58, and 52 by atoms, respectively, when calculated for PSnet= 0. The optimum ratios were inversely related to PSnet and decreased to approximately 20 when PSnet was near maximum. The potential for flowing water to mediate nutrient partitioning among lotic algae by altering growth rates and optimum nutrient ratios is discussed.  相似文献   

19.
De novo synthesis of purine nucleotides and some regulatory properties of this pathway were studied in cultured epithelial-like rat liver cells. It was found that the physiological 5-phosphoribosyl 1-pyrophosphate (P-Rib-PP) concentration in these cells is limiting for purine synthesis de novo. Increase of P-Rib-PP availability, achieved by activation of P-Rib-PP synthetase at high Pi concentration, resulted in acceleration of purine synthesis. The effects of increasing cellular ribose 5-phosphate (Rib-5-P availability, by methylene blue-induced acceleration of the oxidative pentose phosphate pathway, on P-Rib-PP availability and on the rate of the novo purine synthesis were also studied. It was found that at the Pi concentration prevailing in the tissue at extracellular physiological Pi concentration, Rib-5-P availability is saturating for P-Rib-PP generation and therefore also for purine synthesis.  相似文献   

20.
The preincubation of rat liver crude extracts with ATP caused a 60% inactivation of phosphoprotein phosphatase in 30 min at 30 °C. The presence of Mg2+, or cyclic AMP, along with ATP in the preincubation mixture had no effect on the inactivation of phosphatase caused by ATP. The crude liver phosphatase was also inactivated by ADP or PPi; PPi being the most potent inactivating metabolite. AMP, adenosine or Pi were without any effect. The effect of ATP or PPi was completely reversed by cobalt. The cobalt effect was very specific and could not be replaced by several metal ions tested except by Mn2+ which was partly active. With the aid of sucrose density gradient studies, it was also shown that PPicauses an apparent conversion of a 4.1 S form to a 7.8 S form of the enzyme in rat liver extracts. Cobalt, on the other hand, converts the higher 7.8 S form to a lower 4.1 S form of the enzyme. The preincubation of purified rabbit liver phosphoprotein phosphatase with PPi also caused a complete inactivation of the enzyme in 40 min. The inactivation of the enzyme by PPi was completely reversed by cobalt. Unlike the apparent interconversion between different molecular forms of the enzyme by PPi and cobalt in rat liver crude extracts, no such interconversion of purified rabbit liver phosphoprotein phosphatase was observed in the presence of PPi and cobalt.  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