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 effect of trifluoperazine (TFP) on the ATPase activity of soluble and paniculate F1ATPase and on ATP synthesis driven by succinate oxidation in submitochondrial particles from bovine heart was studied at pH 7.4 and 8.8. At the two pH. TFP inhibited ATP hydrolysis. Inorganic phosphate protected against the inhibiting action of TFP. The results on the effect of various concentrations of phosphate in the reversal of the action of TFP on hydrolysis at pH 7.4 and 8.8 showed that H2PO 4 is the species that competes with TFP. The effect of TFP on oxidative phosphorylation was studied at concentrations that do not produce uncoupling or affect the aerobic oxidation of succinate (<15M). TFP inhibited oxidative phosphorylation to a higher extent at pH 8.8 than at pH 7.4; this was through a diminution in theV max, and an increase in theK m for phosphate. Data on phosphate uptake during oxidative phosphorylation at several pH showed that H2PO 4 is the true substrate for oxidative phosphorylation. Thus, in both synthesis and hydrolysis of ATP, TFP and H2PO 4 interact with a common site. However, there is a difference in the sensitivity to TFP of ATP synthesis and hydrolysis; this is more noticeable at pH 8.8, i.e. ATPase activity of soluble F1 remains at about 40% of the activity of the control in a concentration range of TFP of 40–100M, whereas in oxidative phosphorylation 14M TFP produces a 60% inhibition of phosphate uptake.  相似文献   

2.
Enteropeptidase (enterokinase EC 3.4.21.9), catalyzing trypsinogen activation, exhibits unique properties for high efficiency hydrolysis of the polypeptide chain after the N-terminal tetraaspartyl-lysyl sequence. This makes it a convenient tool for the processing of fusion proteins containing this sequence. We found the enteropeptidase-catalysing degradation of some bioactive peptides: cattle hemoglobin beta-chain fragments Hb (2–8) (LTAEEKA) and Hb (1–9) (MLTAEEKAA), human angiotensin II (DRVYIHPF) (AT). Model peptideswith truncated linker WDDRG and WDDKG also were shown to be susceptible to enteropeptidase action. Kinetic parameters ofenteropeptidase hydrolysis for these substrates were determined.K m values for all substrates with truncated linker (10-3 M) are an order of magnitude higher thancorresponding values for typical enteropeptidase artificial peptide or fusion protein substrates with full enteropeptidase linker –DDDDK– (K m 10-4 M). k cat values for AT, Hb (2–8), WDDRG and WDDKG are 30–40 min-1. But one additional amino acid residue at both N- and C-terminus of Hb (2–8) results in a drastic increase of hydrolysis efficiency: k cat value for Hb (1–9) is 1510 min-1. Recent study demonstrates the possibility of undesirable cleavage of target peptides or proteins containing the above-mentioned truncated linker sequences; further, the ability of enteropeptidase to hydrolyse specifically several biologically active peptides in vitro along with its unique natural substrate trypsinogen was demonstrated.  相似文献   

3.
Rate constants have been determined for hydrolysis of the acetate, glutarate, and phthalate monoesters of 2-hydroxy-1,10-phenanthroline in water at 30°C and μ = 0.1 M with KCl. The hydrolysis reactions of the esters are hydroxide ion catalyzed at pH > 9. The phthalate and glutarate monoesters have in addition pH-independent reactions from pH 5.5 to 9 that involve intramolecular participation by the neighboring carboxylate anion. The pH-independent reaction of the glutarate monoester is 5-fold faster than that of the phthalate monoester. The plots of log kobsd vs pH for hydrolysis of the carboxyl substituted esters are bell shaped at pH < 5, which indicates a rapid reaction of the zwitterionic species (carboxyl anion and protonated phenanthroline nitrogen). The divalent metal ions, Cu2+, Ni2+, Zn2+, and Co2+, complex strongly with the esters; saturation occurs at metal ion concentrations less than 0.01 M. The 1:1 metal ion complexes have greatly enhanced rates of hydrolysis; the second-order rate constants for the OH reactions are increased by factors of 105 to 108 by the metal ion. The pH-rate constant profiles for the phthalate and glutarate ester metal ion complexes have a sigmoidal region below pH 6 that can be attributed to a metal ion-promoted carboxylate anion nucleophilic reaction. The carboxyl group reactions are enhanced 102 - to 103 -fold by the metal ions, which allows the neighboring group reaction to be competitive with the favorable metal ion-promoted OH reaction at pH < 6, but not at pH > 6. The half-lives of the pH-independent neighboring carboxyl group reactions of the Cu(II) complexes at 30°C are l2 s. The other metal ion complexes are only slightly less reactive (half-lives vary from 2.5 to 40 s). These are the most rapid neighboring carboxyl group reactions that have been observed in ester hydrolysis.  相似文献   

4.
Summary An amylolytic lactic acid bacterium identified as Lactobacillus plantarum was isolated from cassava roots (Manihot esculenta var. Ngansa) during reting. The amylolytic enzyme synthesized was an extracellular -amylase with an optimum pH of 5.0 and an optimum temperature of 55° C. Cultured on starch, the strain displayed a growth rate of 0.43 h–1, a biomass yield of 0.19 g·g–1 and a lactate yield of 0.81 g·g–1. The growth kinetics were similar on starch and glucose. Sufficient enzyme was synthesized and starch hydrolysis was not a limiting factor for growth. Biosynthesis of the enzyme was observed when the glucose concentration was less than 6.7 g·l–1 and reached up to 4 IU·ml–1 at the end of the fermentation. Offprint requests to: M. Raimbault  相似文献   

5.
Aluminium-tolerant genotypes of Cynodon dactylon are potential candidates for the vegetation of gold mine tailings in South Africa. As a prerequisite to in vitro selection of tolerant genotypes, this work aimed at assessing and adapting micropropagation media to ensure Al3+ activity and toxicity. This was investigated using MINTEQA2, a chemical equilibrium speciation model. The maximum AlAl3+ activity achieved in any medium was 7.5 M. Of the seven published media investigated, four never achieved an activity greater than 4 M at 3–4 mM aluminium. The most appropriate medium was that of Yamamoto et al. (1996) (modified MS without KH2PO4 and EDTA), as it showed an increasing range of AlAl3+ activities from 2 to 7.5 M at aluminium concentrations from 0.25–2.5 mM. An improved modified MS formulation retaining phosphate was investigated because phosphate is an important component of our medium for callus induction in C. dactylon. Using MINTEQA2, no reduction in AlAl3+ activity by phosphate was detected in standard MS medium at pH 4. Through further simulations a new modified MS medium was derived with 1 mM SO 4 2– and no EDTA at pH 4, which gave the maximum AlAl3+ activity (7.5 M) at 2 mM aluminium. This medium gave the highest AlAl3+ activities for the 0.25–2 mM concentration range of all the tested formulations, including the seven published media. It also resulted in significantly higher callus growth rates than standard MS media and other tested media. This new medium is currently being used to screen C. dactylon for aluminium tolerance at pH 4.  相似文献   

6.
Direct electrometry was used to study the light-induced voltage changes in the Rhodobacter sphaeroides chromatophores adsorbed to a phospholipid-impregnated nitrocellulose film. After the second laser flash, a fast increase in the voltage associated with charge separation was followed by a slower increase attributed to the proton uptake in the QB site of the photosynthetic reaction centers. Kinetics and relative amplitudes of these voltage changes attributed to the QA –. B –. QAQBH2 transition, were measured as a function of pH and temperature between +4 and +40 °C. The kinetics can be approximated by a single exponent above +23 °C (100 µs at +25 °C, pH 7.2), whereas below this temperature, it was a good fit of two exponential approximation (65 µs and 360 µs with similar contributions at +10 °C, pH 7.2). The faster component diminished with an apparent pK 8.5, whereas the slower one was maintained at a constant level until pH 9.5 and then decreased. The calculated activation energy from the temperature dependence of the slower component (55 – 65 kJ/mol) was much higher than that of the faster component (< 10 kJ/mol). The two voltage components can be attributed to the transfer of the first (faster component) and the second (slower component) proton from the reaction center surface to QB. We suggested that higher activation energy of the slower component was due to a conformational change in the reaction center kinetically coupled to the second proton transfer to QBH.The faster component diminished in the presence of 1 M KCl, with an apparent pK 7.5. To explain this observation, we assume that: (i) the midpoint potential of the QA/QA –. redox pair was higher in 1 M KCl because of the reduced surface potential of chromatophores; (ii) the midpoint potential of the QB –./QBH–. redox pair was insensitive to the surface potential change; (iii) the equilibrium constant of the reaction QA –.QB –. QAQBH decreased at high ionic strength.  相似文献   

7.
Zusammenfassung Der anAllium-Epidermen unter Deckglas eintretende Asphyxieeffekt zeigt zwischen pH=7,1 und 8,0 eine Abhängigkeit derart, daß er bei niederem pH-Wert rasch (nach 2—4 Minuten), mit steigendem pH-Wert aber immer langsamer (bei pH=8,0 nach 10 Minuten oder später) einsetzt.Phenylmercuriacetat und Jodessigsäure verhindern in Konzentrationen von 10–4 und 10–5 M/l die Anfärbung der Vakuole durch Neutralrot, wie sie auch eine schon vollzogene Anfärbung rückgängig machen.Beide SH-Gruppen-Gifte lösen in gleicher Weise wie Neutralrot Vakuolenkontraktionen aus.Isolierte Vakuolen kollabieren bei Behandlung mit einem der SH-Gruppen-Gifte. Bei geringer Giftkonzentration (10–6 und 10–7 M/l) ist der Volumenschwund langsamer und von granulären Entmischungen begleitet.
Effect of reduced oxygen supply and of inhibitors of SH-groups on the staining of upper epidermal cells ofAllium cepa with neutral red
Summary The influence of pH on the asphyxia-effect in the upper epidermal cells of the second bulb scale of onion has been investigated in time course studies. Under a cover glass (exclusion of air!) neutral red stain shifts from the vacuole to the cell wall within 2–4, 6–9, and over 10 minutes at pH 7.1, 7.5, and at 8.0, or above, respectively.The SH-group inhibitors, phenylmercuric acetate and iodoacetic acid, at concentrations of 10–4 and 10–5 M/l, reduce the staining of the vacuole with neutral red which is accompanied by a corresponding increase in the staining of the wall.The inhibitor-induced contraction of the vacuole is similar to, though relatively weaker than that caused by a basic solution of neutral red.Isolated vacuoles suffer an extreme contraction when treated with the inhibitors at conc. of 10–4 and 10–5 M/l or with per-iodic acid. Weaker concentrations, 10–6 and 10–7 M/l, of the inhibitors bring about a separation of granular bodies followed by a gradual contraction of the vacuoles. Such effects on the isolated vacuoles following treatment with neutral red have been reported earlier (Wartenberg 1961 a).The dye is obviously adsorbed on the free SH-groups and an inhibition of the capacity to accumulate it results from blocking of the SH-groups by the heavy metal compounds.
  相似文献   

8.
Summary The presence of soluble polymeric complexes in dilute solutions (10–30 M) of aluminium and phosphate at 25°C was demonstrated by paper electrophoresis and confirmed by chemical analyses and measurements of 32 P specific activities. Complexed aluminium and phosphate were not determined by colorimetric procedures without preliminary hydrolysis in 1 N HCl at 80–90°C. The formation of aluminophosphate complexes was maximal around pH 5 when, at an initial P/Al mole ratio of 1.6, 3 moles of P were complexed with 4 moles of Al. At pH 6.9, however, only 1 mole of P was removed for every 10 of Al, suggesting that hydroxyaluminium polymers predominated at higher ratios of OH-ions to H2PO4-ions. re]19750523  相似文献   

9.
Enteropeptidase (enterokinase, EC 3.4.21.9) hydrolyzes peptide bonds formed by carboxyl groups of Lys or Arg residue if less than four negatively charged amino acid residues are in positions P 2P 5 of its substrate. We determined the kinetic parameters of three substrates of this type: human angiotensin II (AT) (DR VYIHPF) and the Hb(2–8) (LTAEEK A) and Hb(1–9) (MLTAEEK AA) peptides of the cattle hemoglobin -chain. The K m values for all the substrates (10–3 M) were one order of magnitude higher than those of the typical synthetic substrates of enteropeptidase or chimeric proteins with the –DDDDK– full-size linker (K m 10–4 M). The k cat values for AT and Hb(2–8) were also close and low (30 min–1). The general hydrolysis efficiency of such substrates is no more than 1% of the corresponding value for the typical peptide and protein substrates of the enteropeptidase. However, the elongation of Hb(2–8) peptide by one amino acid residue from both its N- and C-termini results in a dramatic increase in the catalytic efficiency of the hydrolysis: the k cat value for Hb(1–9) is 1510 min–1, which means that it is hydrolyzed only three times less effective than the chimeric protein with the full-size linker.  相似文献   

10.
Summary A strain ofFusarium moniliforme, previously used for microbial protein production, excreted lactase (-D-galactosidase, EC.3.2.1 23) when cultivated either in a whey liquid medium or on a wheat bran solid medium. The enzyme produced in both media had pH and temperature optima of 4–5 and 50–60°C respectively and was particularly suitable for processing acid whey.In the whey culture, maximum lactase yield was observed after 95 h of growth at 30°C and whey lactose concentration of 9%. The addition of ammonium, potassium and sodium ions to the growth medium considerably enhanced lactase production. A maximum enzyme yield corresponding to hydrolysis of 3 nmoles o-nitrophenyl--D-galactopyranoside sec–1 ml–1 of growth medium, at pH 5 and 60°C, was obtained.In the wheat bran culture, the maximum enzyme yield was obtained after 140 h of growth at 28–30°C. A marked increase in the enzyme production was observed when nitrate or phosphate was added to the growth medium. Also, the addition of certain agricultural by-products (molasses, whey) enhanced lactase production. The observed maximum yield corresponding to the hydrolysis of 182 nmoles of ONPG sec–1 g–1 of wheat bran, at pH 5 and 60°C, is comparable to that reported for certain microorganisms used commercially for lactase production.  相似文献   

11.
Summary In the histochemical detection of -D-glucosidase the indigogenic method of Pearson et al. was tested and evaluated. 4-Cl-5-Br-3-indolyl--D-glucoside was used as the substrate.Intestinal -D-glucosidase is firmly bound to the structure. About 60% of activity survives 2 hours fixation in cold 4% formaldehyde and some activity can be demonstrated even in paraffin sections after a shortened embedding.The localization obtained with the original method is not correct. Due to a slow oxidation of indoxyl in the acid pH range and to hydrogen peroxide formation indigo is deposited in sites with an active peroxidase or pseudoperoxidase. The addition of horse-radish peroxidase improves the localization but does not entirely prevent diffusion artifacts. A striking improvement of the localization can be achieved by a mixture of ferri-ferrocyanide. 3.1 · 10–3 M concentration of this oxidation catalyst which is still very effective causes only a 26% inhibition of the enzyme activity as revealed by biochemical assays in homogenates of rat intestine.A new medium was devised consisting of 0.1 M citrate phosphate buffer pH 5.5, 8 · 10–4 M 4-Cl-5-Br-3-indolyl--D-glucoside and 3.1 · 10–3 M ferri-ferrocyanide mixture. With this medium a very clear brush border localization of the enzyme activity (activities) in enterocytes of the rat and human intestine was demonstrated. This activity is present in differentiated enterocytes covering the villi. The highest activity resides in the jejunum. Enzyme activity is considerably inhibited by galactonolactone (5 · 10–3 M) and gluconolactone (4 · 10–4 M). It is completely inhibited by florizin (5 · 10–3 M). Cellobiose (8 · 10–2 M) caused 65%, lactose (8 · 10–2 M) 48% and glucose (8 · 10–2 M) 35% inhibition (data were obtained by cytospectrophotometry). In patients with celiac sprue the activity is very much decreased. Its restitution after a gluten-free diet proceeds roughly parallel to that of lactase. The relationship of the demonstrated activity (activities) to florizin hydrolase and lactase is discussed.In the kidney the reaction is very weak and is confined to the cells of proximal convoluted tubules (diffuse staining with some enhancement at the luminal surface of the brush border). The method is also very useful for processing zymograms.  相似文献   

12.
The medicinal mushroom Agaricus blazei produced high amounts of laccase (up to 5,000 units l–1) in a complex, agitated liquid medium based on tomato juice, while only traces of the enzyme (<100 units l–1) were detected in synthetic glucose-based medium. Purification of the enzyme required three chromatographic steps, including anion and cation exchanging. A. blazei laccase was expressed as a single protein with a molecular mass of 66 kDa and an isoelectric point of 4.0. Spectroscopic analysis of the purified enzyme confirmed that it belongs to the blue copper oxidases. The enzymes pH optimum for 2,6-dimethoxyphenol (DMP) and syringaldazine was pH 5.5; but for 2,2-azino-bis(3-ethylthiazoline-6-sulfonate) (ABTS) no distinct pH optimum was observed (highest activity at the lowest pH tested). Purified laccase was stable at 20°C, pH 7.0 and pH 3.0, but rapidly lost its activity at 40°C or pH 10. Sodium chloride strongly inhibited the enzyme activity, although the inhibition was completely reversible. The following kinetic constants were determined (Km, kcat): 63 M, 21 s–1 for ABTS, 4 M, 5 s–1 for syringaldazine, 1,026 M, 15 s–1 for DMP and 4307 M, 159 s–1 for guaiacol. The results show that—in addition to the wood-colonizing white-rot fungi—the typical litter-decomposing basidiomycetes can also produce high titers of laccase in suitable liquid media.  相似文献   

13.
D. Sanders  F. A. Smith  N. A. Walker 《Planta》1985,163(3):411-418
Rapid lowering of the external pH (pH jump) enhances Cl influx in Chara. Experiments were conducted to distinguish between two factors which have previously been proposed to mediate in the response: raised cytoplasmic pH and lowered cytoplasmic Cl concentration. It is concluded that the latter alternative is more likely because: i) Cl influx is reduced at high external pH; ii) influx following the pH jump is never greater than that following pretreatment in Cl-free solution, which reduces cytoplasmic Cl concentration (Cl starvation); iii) the joint application of pH jump and Cl starvation does not result in a greater Cl influx than does Cl starvation alone; and iv) addition of NH 4 + , which increases cytoplasmic pH, does generate an additional stimulation of Cl influx following a pH jump. It is suggested that the increased cytoplasmic pH at the end of pretreatment at high external pH decays rapidly during the pH jump, and thus any effect on Cl influx is so transient as to be undetectable by the methods used. The results are discussed in terms of a reaction kinetic model for 2H+/Cl cotransport (Sanders, D. and Hansen, U.-P, 1981, J. Member. Biol. 58, 139–153) which describes quantitatively; i) the effects of NH 4 + on Cl influx in terms involving only a change in cytoplasmic pH; and ii) the combined effects of Cl starvation and NH 4 + in terms involving only changes in Cl concentration and cytoplasmic pH. Conversely, the combined effects of Cl starvation and pH jump cannot be described by the model if the effect of the pH jump is the consequence of increased cytoplasmic pH. The simple interpretation of experiments on whole cells involving manipulation of (the electrochemical potential difference for protons across the plasma membrane) is questioned in the light of these and previous findings that secondary factors can determine the response of Cl transport in Chara.Abbreviations CPW Chara pond water - [Cl]c cytoplasmic Cl concentration - pHc cytoplasmic pH - pHo external pH - transmembrane electrochemical gradient of protons - a membrane electrical potential difference  相似文献   

14.
Summary Soft water of low buffer capacity was drawn from near the branchial surface of rainbow trout (Salmo gairdneri) at 15°C, using opercular catheters, to determine pH changes in water passing over the gills. Latex masks allowed measurement of ventilation volume, and concentrations of carbon dioxide, oxygen, ammonia, and titratable base in expired water were compared to concentrations in inspired water. Water passing over the gills was more basic than inspired water if the inspired water was pH 4–6 (maximum increase: +0.7 pH units near pH 5). Expired water was more acidic than inspired water if the inspired water was pH 6–10 (maximum decrease: –1.7 pH units near pH 9). Ventilation volume (0.37 l·kg–1·min–1) and oxygen consumption (1.7 mmol·kg–1·h–1) were constant in the pH range 4.6–10.1, but both increased by 1.6–2.4× near pH 4. Carbon dioxide transfer near the gills was about 100 M, ammonia transfer about 15 M, and titratable base added at the gills was about 30 M. A theoretical model using CO2, titratable base, and ammonia added at the gills, the titration characteristics of the defined soft water medium, and aquatic equilibria for CO2 and ammonia, adequately explained the experimentally observed changes in pH near trout gills. Our observations and predictive model indicate that any gill contaminant whose toxicity varies with pH may be more or less toxic at the gills than predicted from bulk water chemistry alone.Abbreviations pH ex expired pH - pH in inspired pH  相似文献   

15.
In oligotrophic waters, not only community structure but also physiological properties of heterotrophic bacteria are influenced by the concentration of organic matter.The relationship between growth rate of two facultatively oligotrophic strains ofAeromonas sp. No. 6 andFlavobacterium sp. M1 was studied in comparison with that of two eutrophic strains ofEscherichia coli 7020 andFlavobacterium sp. M2. These strains had two or three different substrate constants (Ks values) depending on substrate concentrations: Ks values for the two former were remarkably lower than those for the two latter. For instance, Ks value forAeromonas sp. No. 6 was about 8.9M when substrate concentration was greater than 53M and about 1.1M when substrate concentration was less man 53M. InE. coli the Ks value was about 260M at greater than 5600M and about 47M at less than 5600M substrate concentration.Uptake kinetics ofAeromonas sp. grown in a medium containing 2.7 mM glutamate (H-cell) and 0.11M glutamate (L-cell) have been determined for the intact cells. H-cell had two distinct values of Km for glutamate assimilation and respiration, and L-cell had three distinct values of Km for glutamate assimilation and respiration: In H-cell Km of assimilation was 2.8×10–7 M and 1.5×10–4 M, and Km of respiration was 2.3×10–7 M and 1.7×10–4 M; in L-cell Km of assimilation was 7.4×10–8 M, 8.3×10–6 M, and 1.3×10–4 M, and Km of respiration was 2.5×10–7, 8.9×10–6M, and 1.7×10–4 M. More than 60% of glutamate taken up by the H- and L-cells was respired when the substrate concentration was less than 10–6 M, although at greater than 10–6 M, 50% and 30% of glutamate was respired by H-cells and L-cells, respectively. These results suggest that the facultatively oligotrophic bacteria grow with high efficiency in environments with extremely low nutrient concentration, such as oligotrophic waters of lakes and ocean, as compared with in their growth in conditions of high nutrient concentraton, such as nutrient broth.  相似文献   

16.
Peroxidase-catalyzed oxidation of o-phenylenediamine (PDA) is greatly activated with melamine (MA) in 15 mM phosphate–citrate buffer at pH 6.0–7.4 in a noncompetitive manner: k cat and K m increase in direct proportion to the MA concentration. An extent of the activation is quantitatively characterized with a coefficient (in M–1), which essentially increases along with the rise in pH from 6.0 to 7.4. MA acts as a nucleophilic catalyst in the oxidation process: it most likely affects the peroxidase active site from the distal position of heme. MA noncompetitively inhibits the peroxidase oxidation of PDA at pH 4.3, since it completely loses its nucleophilic properties in acidic medium. A rapid, highly accurate, and simple analytical test system based on the kinetics of melamine-activated oxidation of PDA is proposed for the quantitative determination of melamine within the concentration range of 10–4–10–3 M. This test system uses the spectrophotometric determination of the PDA oxidation product at 455 nm.  相似文献   

17.
The interaction of various hormones and regucalcin on (Ca2+–Mg2+)-ATPase activity in rat liver plasma membranes was investigated. The presence of epinephrine (10–6–10–4 M), and insulin (10–8–10 M) in the reaction mixture produced a significant increase in (Ca2+–Mg2+)-ATPase activity, while the enzyme activity was decreased significantly by calcitonin, (3×10–8–3×10–6 M). These hormonal effects, except for calcitonin, were clearly inhibited by the presence of vanadate (10–4 M) which can inhibit the Ca2+-dependent phosphorylation of enzyme. Meanwhile, regucalcin (0.25 and 0.50 M), isolated from rat liver cytosol, elevated significantly (Ca2+–Mg2+)-ATPase activity in the plasma membranes, although this elevation was not inhibited by vanadate (10–4 M). the epinephrine (10–5 M) or phenylephrine (10–4 M)-induced increase in (Ca2+–Mg2+)-ATPase activity was disappeared in the presence of regucalcin; in this case the effect of regucalcin was also weakened. However, the inhibitory effect of calcitonin (3×10–6 M) was not weakened by the presence of regucalcin (0.5 M). Moreover, GTP (10–5 and 10–4 M)-induced increase in (Ca2+–Mg2+)-ATPase activity was not seen in the presence of regucalcin (0.25 M). The present finding suggests that the activating mechanism of regucalcin on (Ca2+–Mg2+)-ATPase is not involved on GTP-binding protein which modulates the receptor-mediated hormonal effect in rat liver plasma membranes.  相似文献   

18.
The effects of eight microelements (I, BO3 3–, MoO4 2–, Co2+, Cu2+, Mn2+, Fe2+, Zn2+) on the biosynthesis of camptothecin and the growth of suspension cultures of Camptotheca acuminata were studied. The increase of I to 25 M l–1, Cu2+ to 1 M l–1, Co2+ to 2 M l–1 and MoO4 2– to 10 M l–1 in Murashige and Skoog (MS) medium resulted in 1.66, 2.84, 2.53 and 2.04 times higher of camptothecin yield than that in standard MS medium respectively. Combined treatment of I (25 M l–1), Cu2+ (1 M l–1), Co2+ (2 M l–1) and MoO4 2– (10 M l–1) lead to improve cell dry weight, camptothecin content, and camptothecin yield to 30.56 g l–1, 0.0299%, and 9.15 mg l–1, respectively, which were 20.2, 208.9 and 273.8% increment respectively when compared with those of control.  相似文献   

19.
Summary Relationships between root zone temperature, concentrations and uptake rates of NH 4 + and NO 3 were studied in non-mycorrhizal roots of 4-year-old Norway spruce under controlled environmental conditions. Additionally, in a forest stand NH 4 + and NO 3 uptake rates along the root axis and changes in the rhizosphere pH were measured. In the concentration (Cmin) range of 100–150 M uptake rates of NH 4 + were 3–4 times higher than those of NO 3 The preference for NH 4 + uptake was also reflected in the minimum concentration (Cmin) values. Supplying NH4NO3, the rate of NO 3 uptake was very low until the NH 4 + concentrations had fallen below about 100 M. The shift from NH 4 + to NO 3 uptake was correlated with a corresponding shift from net H+ production to net H+ consumption in the external solution. The uptake rates of NH 4 + were correlated with equimolar net production of H+. With NO 3 nutrition net consumption of H+ was approximately twice as high as uptake rates of NO 3 In the forest stand the NO 3 concentration in the soil solution was more than 10 times higher than the NH 4 + concentration (<100 M), and the rhizosphere pH of non-mycorrhizal roots considerably higher than the bulk soil pH. The rhizosphere pH increase was particularly evident in apical root zones where the rates of water and NO 3 uptake and nitrate reductase activity were also higher. The results are summarized in a model of water and nutrient transport to, and uptake by, non-mycorrhizal roots of Norway spruce in a forest stand. Model calculations indicate that delivery to the roots by mass flow may meet most of the plant demand of nitrogen and calcium, and that non-mycorrhizal root tips have the potential to take up most of the delivered nitrate and calcium.  相似文献   

20.
Summary A stopped-flow rapid reaction apparatus was used to study the rate of pH equilibration in human red cell suspensions. Flux of OH or H+ was determined over a wide range of extracellular pH (4–11) in CO2-free erythrocyte suspensions. In these experiments, an erythrocyte suspension at pH 7.3 is rapidly mixed with an equal volume of NaCl solution at 3.0>pH>11.5. The pH of the extracellular fluid of the mixture changes rapidly as OH or H+ moves across the red cell membrane. Flux and velocity constants can be calculated from the initiald pH/dt using the known initial intra- and extracellular pH. It was found that the further the extracellular pH is from 7.3 (in either direction from 4–11), the greater the absolute value of total OH and/or H+ flux. Pretreatment with SITS (4-acetamido-4-isothiocyanostilbene-2,2-disulfonic acid), a potent anion exchange inhibitor, greatly reduces flux over the entire pH range, while exposure to valinomycin, a potassium ionophore, has no measurable effect. These data suggest that (i) both H+ and OH may be moving across the red cell membrane at all pH; (ii) the species dominating pH equilibration is probably dependent on the extracellular pH, which determines the magnitude of the driving gradient for each ion; and (iii) the rapid exchange pathway of the erythrocyte membrane may be utilized for both H+ and OH transport during CO2-free pH equilibration.  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