首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The effect of nitrogen source concentration on the production of the polysaccharide curdlan by the bacterium Agrobacterium sp. ATCC 31749 from hydrolysates of prairie cordgrass was examined. The highest curdlan concentrations were produced by ATCC 31749 when grown on a medium containing a solids-only hydrolysate and the nitrogen source ammonium phosphate (2.2 mM) or on a medium containing a complete hydrolysate and 3.3 mM ammonium phosphate. The latter medium sustained a higher level of bacterial curdlan production than the former medium after 144 hr. Biomass production by ATCC 31749 was highest after 144 hr when grown on a medium containing a solids-only hydrolysate and 2.2 or 8.7 mM ammonium phosphate. On the medium containing the complete hydrolysate, biomass production by ATCC 31749 was highest after 144 hr when 3.3 mM ammonium phosphate was present. Bacterial biomass production after 144 hr was greater on the complete hydrolysate medium compared to the solids-only hydrolysate medium. Curdlan yield produced by ATCC 31749 after 144 hr from the complete hydrolysate medium containing 3.3 mM ammonium phosphate was higher than from the solids-only hydrolysate medium containing 2.2 mM ammonium phosphate.  相似文献   

2.
Abstract: Microdialysis was used in a comparative study of the neurotoxic action of MPP+ in the absence or presence of nomifensine (20 µM) in the striatum and substantia nigra. Three different concentrations of MPP+ (1, 2.5, and 5 mM) were perfused for 15 min at 24 (day 1) and 48 h (day 2) after surgery. The dopamine basal value in the striatum was ~17 fmol/min. Nomifensine (20 µM) stimulated dopamine release to ~170 fmol/min. The increase of dopamine extracellular output in the striatum after MPP+ perfusion on day 1 was independent of the concentration of MPP+ perfused and of the absence or presence of nomifensine (20 µM), being ~2,500 fmol/min. The dopamine basal value in the substantia nigra was below the detection limit of our HPLC equipment. Nomifensine (20 µM) stimulated dopamine release to ~6.3 fmol/min. The increase of dopamine extracellular output in the substantia nigra was MPP+ dose-dependent (1 mM, 75 fmol/min; 2.5 mM, 150 fmol/min; and 5 mM, 250 fmol/min) and independent of the presence or absence of nomifensine. On day 2, the presence of nomifensine on day 1 produced a total protection against MPP+ (1 mM) perfusion in the striatum, which was not observed against MPP+ (5 mM). MPP+ (1 mM) did not produce any neurotoxic action in the substantia in the absence or presence of nomifensine. The MPP+ (2.5 mM) effect on dopamine extracellular output in the absence of nomifensine (20 µM) in the substantia nigra on day 2 was similar to that of MPP+ (1 mM) in the striatum. The presence of nomifensine (20 µM) partially prevented the neurotoxic effect of MPP+ (2.5 mM) on dopaminergic cell bodies/dendrites in the substantia nigra. The MPP+ (5 mM) effect on dopamine extracellular output was similar in both structures studied in the absence or presence of nomifensine on day 2. These results suggest that terminals in the striatum are more sensitive to the neurotoxicity of MPP+ than cell bodies/dendrites in the substantia nigra.  相似文献   

3.
Summary Actinorhodin production inStreptomyces coelicolor A3(2) was relatively insensitive to the carbon source concentration but was elicited by nitrogen or phosphate depletion, or by a decline in the growth rate. In starch-glutamate media with nitrogen limitation, increasing the nitrogen supply delayed the onset of antibiotic synthesis and, at concentrations above 30 mM, decreased its rate. In a similar medium with phosphate limitation, increasing the initial phosphate concentration delayed actinorhodin formation and, above 2.5 mM, reduced the rate of synthesis. Experiments in which actinorhodin synthesis was elicited by phosphate depletion at various nitrogen concentrations demonstrated strong suppression by residual glutamate. Cultures in which actinorhodin biosynthesis was initiated by nitrogen depletion were not similarly suppressed by increasing amounts of residual phosphate. The results suggest that actinorhodin production inS. coelicolor A3(2) responds to interacting physiological controls, notable among which is nitrogen catabolite regulation.  相似文献   

4.
Modulation of the sinus rate and contractile force by taurine at different extracellular Ca2+ concentrations ([Ca2+]o) was examined using rat right atria loaded with forced swimming stress. Serum concentration of corticosterone profoundly increased in stress-loaded rats as compared with native rats. The taurine level in serum also increased in stress-loaded rats, but was not changed in the different heart tissues and aorta. Heat-shock protein (HSP72) was detectable in cardiac muscles and in the lumen of cardiac blood vessels of stress-loaded rats using a monoclonal antibody. Increasing [Ca2+]o (from 0.9 to 3.6 mM) enhanced the sinus rate and contractile force in a [Ca2+]o-dependent fashion in native rats, but not in stress-loaded rats. Taurine (1–20 mM) caused a negative chronotropic and inotropic effect in a concentration-dependent manner. At 1.8 mM [Ca2+]o, the negative chronotropic effect of taurine (10–20 mM) was attenuated in stress-loaded rats as compared with native rats. These results indicate that swimming stress causes a release of taurine into the serum and reduces the sensitivity to [Ca2+]o. Taurine administration might, in part, exhibit the protective actions on acute stress-induced responses.  相似文献   

5.
The objectives of this study were to examine the effects of growth substrate and extracellular pH on phosphoenolpyruvate-dependent glucose phosphorylation as well as to examine how maltose is phosphorylated by the ruminal bacterium Megasphaera elsdenii B159. Phosphoenolpyruvate-dependent glucose phosphorylation by toluene-treated cells was constitutive, and glucose phosphorylation was reduced by 69% at pH 5.0. When toluene-treated cells were incubated in histidine buffer, little maltose phosphorylation occurred in the absence of inorganic phosphate. However, the addition of increasing concentrations of either potassium or sodium phosphate increased maltose phosphorylation. Maximal phosphorylation activity was observed at between 25 and 50 mM of either inorganic phosphate source. Compared with the control incubations, maltose phosphorylation was increased over threefold with 25 mM of either potassium or sodium phosphate. Phosphoglucomutase activity was detected in cell extracts of M. elsdenii B159, and this enzyme had a K m of 3.2 mM for glucose-1-P and a V max of 1836 nmol of NADP+ reduced/mg of protein per min. Maltose was also hydrolyzed by an inducible maltase (K m , 1.19 mM). To our knowledge, this is the first report of a maltose phosphorylase and a maltase in M. elsdenii. Received: 3 November 1999 / Accepted: 5 January 2000  相似文献   

6.
An NADP-specific glutamate dehydrogenase [L-glutamate: NADP+ oxidoreductase (deaminating), EC 1.4.1.4] from alkaliphilic Bacillus sp. KSM-635 was purified 5840-fold to homogeneity by a several-step procedure involving Red-Toyopearl affinity chromatography. The native protein, with an isoelectric point of pH 4.87, had a molecular mass of approximately 315 kDa consisting of six identical summits each with a molecular mass of 52 kDa. The pH optima for the aminating and deaminating reactions were 7.5 and 8.5, respectively. The optimum temperature was around 60°C for both. The purified enzyme had a specific activity of 416units/mg protein for the aminating reaction, being over 20-fold greater than that for deaminating reaction, at the respective pH optima and at 30°C. The enzyme was specific for NADPH (Km 44 μM), 2-oxoglutarate (Km 3.13 mM), NADP+ (Km 29 μM), and L-glutamate (Km 6.06 mM). The Km for NH4Cl was 5.96 mM. The enzyme could be stored without appreciable loss of enzyme activity at 5°C for half a year in phosphate buffer (pH 7.0) containing 2 mM 2-mercaptoethanol, although the enzyme activity was abolished within 20 h by freezing at ?20°C.  相似文献   

7.
l-Serine dehydratase fromLactobacillus fermentum was purified 100-fold. It was stabilized by the presence of 1 mM l-cysteine in 50 mM phosphate buffer. Mr=150,000 was determined by gel filtration. The enzyme consists of four apparently identical subunits (Mr=40,000) that were observed after treatment with sodium dodecyl sulfate. The apparent Km forl-serine was 65 mM. Fe++ was required for the enzymatic activity, and the apparent Km value for this reaction was 0.55 mM. Maximum enzymatic activity was observed at 45°C and pH 8.0 in 50 mM phosphate buffer. At pH values different from the optimum, a positive cooperativity between substrate molecules was observed. The activation energy of the reaction was 11,400 and 22,800 cal × mol–1 for temperature values more than and less than 35°C respectively. The purified enzyme showed a maximum absorption between 400 and 420 nm, indicating the presence of pyridoxal-5-phosphate (PLP) as a prosthetic group. The PLP concentration was 0.027 µmoles per milligram of protein. The data suggest that there is 1 mol of PLP for each protein subunit.  相似文献   

8.
Summary The influence of inorganic phosphate and immobilization on cells of Claviceps purpurea strain 1029/N5 producing ergot peptides in shake culture was examined. Immobilization in Ca-alginate beads resulted in a marked reduction of some metabolic activities, i.e. the periods of alkaloid formation and cell growth were prolonged. High concentrations of inorganic phosphate (1 g/l KH2PO4) could reduce or stop alkaloid formation both by free and immobilized cells at any time during fermentation. The optimum phosphate concentration for alkaloid production by immobilized cells (about 0.5 mM) was a quarter of that required by free cells. This optimum shift was attributed to (i) the diminished phosphate demand of immobilized cells, due to their reduced metabolic activities, and (ii) the phosphate-dependent morphological behaviour of the biocatalyst. The observed decrease in alkaloid concentrations during later periods of the fermentation supported the idea of alkaloid-degradative enzymes, activated by high phosphate concentrations. Immobilization showed an advantageous influence on this undesirable effect. Offprint requests to: H.J. Rehm  相似文献   

9.
The effects of cadmium on isolated corn shoot mitochondria were determined. In the absence of phosphate cadmium stimulated the oxidation of exogenous NADH optimally at 0.025 mM, but was inhibitory at 0.1 mM and above. The presence of phosphate negated the cadmium stimulation of exogenous NADH oxidation and permitted inhibitions only at higher cadmium concentrations. Succinate or malate + pyruvate oxidation in the absence of phosphate was inhibited to a greater extent by cadmium than when phosphate was present. ADP/O and respiratory control ratios were reduced by cadmium but generally were less sensitive to cadmium than state 4 or minus phosphate respiration. The data suggest that the site of cadmium effect is likely to be early in electron transport. Cadmium had a pronounced effect on mitochondrial swelling under either passive or active conditions. When succinate or exogenous NADH were being oxidized swelling occurred at 0.05 mM cadmium, but with malate + pyruvate the cadmium concentration had to exceed 1.0 mM. Phosphate (2 mM) prevented the swelling. Dithiothreitol, a SH group protector, prevented any effect of cadmium on swelling or respiration which suggests that sulfhydryl groups are likely involved in the cadmium-membrane interaction.  相似文献   

10.
Proteins from the fresh roots of Stemona tuberosa (Stemonaceae) were extracted into 20 mM phosphate buffer, pH 7.2/0.1 M NaCl, precipitated with 90% saturation ammonium sulfate, and enriched by diethylaminoethanol (DEAE) cellulose. The protein eluted as a single main peak from the unbound fractions (ST-1), and appeared as a single band with superoxide dismutase (SOD) activity after native polyacrylamide gel electrophoresis (PAGE) resolution and zymogram development. ST-1 was classified as SOD due to its strong inhibition by HCN and H2O2. The amino acid sequence of three tryptic peptides of ST-1 matched with the SOD isozymes from Ananas comosus and Solanum lycopersicum. The SOD consisted of at least two heterologous protein subunits with molecular mass of 17.6 and 31.5 kD, respectively, and had an optimal SOD activity at pH 5 and over a temperature range of 0–50°C. MgCl2, MnCl2, and HgCl2 were strongly inhibitory at all concentrations tested. The SOD activity was completely negated in the presence of 0.5 mM SDS or 5 mM HgCl2. The relationship between riboflavin and nitroblue tetrazolium (NBT) on SOD activity was linear, giving K m and V max values of the purified SOD of 62.414 ± 0.015 M and 101.010 ± 0.022 µmol/min/mg protein for NBT and 27.389 ± 0.032 M and 38.167 ± 0.021 µmol/min/mg protein for riboflavin, respectively.  相似文献   

11.
Summary A higher producer of ascospores and pigments, Monascus strain TTWMB 6042, was used to study regulation of pigment production by nutrients. An initial medium containing 4% glucose, 0.3% NH4NO3 (75 mm nitrogen) and inorganic salts was used. We found that the formation of red pigments in this strain, measured by optical density at 500 nm (OD500) was strongly stimulated by monosodium glutamate (MSG) as the sole nitrogen source. The choice of carbon source and an initial pH of pH 5.5 were also important. High concentrations of phosphate and MgSO4 were inhibitory to pigment production. A new chemically defined medium was devised containing 5% maltose, 75 mm MSG, phosphate and MgSO4 at lower concentrations plus other mineral salts, which yielded a tenfold increase in OD500 and a reversal of the pigment location from predominantly cell-bound, including both intracellular and surface-bound pigments, to mainly extracellular. Offsprint requests to: A. L. Demain  相似文献   

12.
Summary The effects of increasing concentrations of NaCl and CaCl2 on quince (Cydonia oblonga Mill. BA 29 clone) somatic embryogenesis and adventitious root regeneration were investigated. Leaves collected from in vitro-grown shoots were used as explants and induced for 2d in liquid Murashige and Skoog medium containing 11.3 μM 2,4-dichlorophenoxyacetic acid. Explants were then cultured on semisolid Murashige and Skoog medium enriched with 4.7 μM kinetin and 0.5 μM naphthaleneacetic acid under red light for 25 d and under white light for another 25 d. Two experiments were performed: in the first, NaCl was used at 0,25, 50, 100, and 200 mM in factorial combination with CaCl2 at 3, 9, and 27 mM; in the second, NaCl was applied at 0, 5, 10, 20, 40, and 80 mM in combination with CaCl2 at 0.3, 1.0, and 3.0 mM. Quince leaves revealed the capacity to regenerate somatic embryos and/or adventitious roots. Quantitative and qualitative regeneration from leaves was affected by NaCl treatments: increasing NaCl concentrations, in combination with CaCl2 at 1 mM, led to an increase in the proportion of leaves producing somatic embryos only, and to a decrease of both leaves regenerating roots only and leaves simultaneously producing somatic embryos and adventitious roots. This suggests a beneficial effect of salt stress on the embryogenic process. The regeneration response decreased with increasing salt concentrations and was almost totally inhibited above 50 mM NaCl and 9 mM CaCl2. The presence of CaCl2 in the culture medium apparently mitigated the effects of salt stress, but only when NaCl was applied at 40 mM. NaCl at 5 mM, in the presence of 0.3 or 1 mM CaCl2, was favorable both to somatic embryo and root production. No value of the ratio Na+/Ca2+ was found to be optimal for the regeneration processes.  相似文献   

13.
2-Deoxyribose 5-phosphate production through coupling of the alcoholic fermentation system of baker’s yeast and deoxyriboaldolase-expressing Escherichia coli was investigated. In this process, baker’s yeast generates fructose 1,6-diphosphate from glucose and inorganic phosphate, and then the E. coli convert the fructose 1,6-diphosphate into 2-deoxyribose 5-phosphate via D-glyceraldehyde 3-phosphate. Under the optimized conditions with toluene-treated yeast cells, 356 mM (121 g/l) fructose 1,6-diphosphate was produced from 1,111 mM glucose and 750 mM potassium phosphate buffer (pH 6.4) with a catalytic amount of AMP, and the reaction supernatant containing the fructose 1,6-diphosphate was used directly as substrate for 2-deoxyribose 5-phosphate production with the E. coli cells. With 178 mM enzymatically prepared fructose 1,6-diphosphate and 400 mM acetaldehyde as substrates, 246 mM (52.6 g/l) 2-deoxyribose 5-phosphate was produced. The molar yield of 2-deoxyribose 5-phosphate as to glucose through the total two step reaction was 22.1%. The 2-deoxyribose 5-phosphate produced was converted to 2-deoxyribose with a molar yield of 85% through endogenous or exogenous phosphatase activity.  相似文献   

14.
For mimicking macromolecular crowding of DNA quadruplexes, various crowding agents have been used, typically PEG, with quadruplexes of micromolar strand concentrations. Thermal and thermodynamic stabilities of these quadruplexes increased with the concentration of the agents, the rise depended on the crowder used. A different phenomenon was observed, and is presented in this article, when the crowder was the quadruplex itself. With DNA strand concentrations ranging from 3 µM to 9 mM, the thermostability did not change up to ~2 mM, above which it increased, indicating that the unfolding quadruplex units were not monomolecular above ~2 mM. The results are explained by self‐association of the G‐quadruplexes above this concentration. The ΔGo37 values, evaluated only below 2 mM, did not become more negative, as with the non‐DNA crowders, instead, slightly increased. Folding topology changed from antiparallel to hybrid above 2 mM, and then to parallel quadruplexes at high, 6–9 mM strand concentrations. In this range, the concentration of the DNA phosphate anions approached the concentration of the K+ counterions used. Volume exclusion is assumed to promote the topological changes of quadruplexes toward the parallel, and the decreased screening of anions could affect their stability. © 2013 Wiley Periodicals, Inc. Biopolymers 101: 428–438, 2014.  相似文献   

15.
The transport of inorganic phosphate has been studied inAcinetobacter lwoffi JW11. During growth on excess phosphate, only one transport system was present, with an apparent Km of 1.4 M. When cells were starved for phosphate, a second uptake system with an apparent Km of 110 nM was also synthesized. The two transport systems could be distinguished by differing sensitivities to the phosphate analogs arsenate and 2-aminoethylphosphonate. Both systems were inhibited by carbonylcyanidem-chlorophenylhydrazone, and to a lesser extent by Na azide. The high-affinity transport system was inactivated by osmotic shock treatment and by spheroplast formation. Preliminary evidence for a phosphate-binding protein in the osmotic shock fluid is presented. The isolation of a mutant constitutive for the high-affinity transport system is described.  相似文献   

16.
Effects of exogenous glycinebetaine (GB, 2–50 mM) on growth, photosynthetic gas exchange, PSII photochemistry, and the activities of key enzymes involved in CO2 fixation in maize plants were investigated. Growth, CO2 assimilation rate, and stomatal conductance increased at low GB concentrations (2–20 mM) but decreased significantly at high GB concentrations (30–50 mM). Leaf relative water content and water potential remained unchanged at low GB concentrations but decreased at high GB concentrations. The maximal efficiency of PSII photochemistry was unchanged either at low or high GB concentrations. The actual PSII efficiency ( Φ PSII) and photochemical quenching (qP) increased at low GB concentrations but decreased at high GB concentrations. At low GB concentrations, there were no significant changes in the efficiency of excitation energy capture by open PSII reaction centres (Fv′/Fm′) and non‐photochemical quenching (qN). At high GB concentrations, Fv′/Fm′ decreased while qN increased significantly. There were no changes in the activities of phosphoenolpyruvate carboxylase, pyruvate phosphate dikinase, and ribulose‐1,5‐bisphosphate carboxylase in control and GB‐fed plants. However, there was a linear correlation between CO2 assimilation rate and stomatal conductance in control and GB‐fed plants. Moreover, there were no significant differences in O2 evolution rate between control and GB fed‐plants under saturated CO2 conditions. The results suggest that exogenous GB application at certain concentrations can enhance CO2 assimilation rate, which can be explained by an increased stomatal conductance.  相似文献   

17.
We report electric-dichroism and electron-microscopic studies of chromatin fibers fixed by protein–protein crosslinking at salt concentrations ranging from 10 to 100 mM. The results confirm a progressive disorganization of the fiber as the salt concentration is lowered. The positive dichroism and large polarizability anisotropy characteristic of the 300-Å diameter fiber found in 100 mM salt are replaced by negative dichroism and smaller effective polarizability anisotropy or dipole moment for samples fixed at lower salt concentration. We interpret the results in terms of segmental, field-induced orientation of the disorganized structure which is present in low salt concentrations. We also observed a field-induced absorbance decrease in chromatin fibers fixed at salt concentration at and below 100 mM. All three optical effects, namely overall orientation of the high-salt fixed fiber, segmental orientation of the low-salt fixed fiber, and field-induced absorbance decrease, occur on roughly the same time scale, 20–100 μs for 50 nucleosome polynucleosomes. The polarizability anisotropy of fibers fixed in 100 mM salt was found to be proportional to the length of the fragment and to the reciprocal square root of the conductivity of the solution used for electric-dichroism measurements. Addition of Mg2+ to the measurement buffer affected the dichroism amplitude of samples fixed below 100 mM salt but not those fixed at 100 mM salt. The results reinforce the need for caution in interpreting electric-dichroism measurements on chromatin fibers because of possible field-induced distortion effects.  相似文献   

18.
Barraclough  P. B. 《Plant and Soil》1993,155(1):175-178
Macronutrients accumulate as free ions in all plant organs when supplies are abundant but growth requirements are generally satisfied before high concentrations are attained. Low threshold concentrations are necessary to ensure maximum growth and in wheat leaves these are around 10 mM nitrate, 5 mM phosphate, 0.5 mM sulfate and 130 mM potassium. The measurement of inorganic ion concentrations in leaves has potential as a diagnostic toll and might enable soil nutrient supplies to be more accurately matched with crop needs.  相似文献   

19.
Earlier studies described the linkage of silver to antibodies using SH groups generated by the reduction of the SS groups using ascorbic acid (1) analogous to the Thakur and DeFulvio technique for linking technetium to antibodies. This work describes the linkage of silver to IgG after introducing SH groups by coupling the IgG to 2-imino thiolane. The protein was dissolved in sodium acetate buffer pH 4.5 containing 1 mM EDTA by dialysis/gel chromatography in a concentration of 20 mg/mL. 2-Imino thiolane dissolved in Tris-HCl acetate buffer, pH 8.2, 0.2M was added to give a final dilution of 0.2 mM 2-imino thiolane. The excess of 2-imino thiolane was removed by dialysis or G-25 Sephadex gel chromatography and then the protein was reacted with silver nitrate 0.1 mM. The unreacted SH groups were blocked by adding iodoacetamide to a concentration of 5 mM. The nonprotein reagents again were removed by dialysis or gel chromatography. The thiol groups were titrated using 1.5 mM 2 2-Py-SS-Py prior to and after addition of silver. It was observed that depending on the concentration of silver, 50–80% of the SH groups were coupled to silver. Higher concentrations of silver led to insoluble precipitates and should be avoided.  相似文献   

20.
Ion Uptake Efficiency of Sunflower Roots   总被引:1,自引:0,他引:1  
The term ion uptake efficiency is used for the rate of uptake of a particular ion from nutrient solutions holding a standard concentration of that ion (0.5 mM sulphate, 1.5 mM phosphate or 2.0 mM rubidium). The uptake efficiency for rubidium and phosphate in roots of intact sunflower plants depended on the salt status of the plants and on the concentration of the ion under investigation in pretreatment solutions. The effect of pretreatment was a rapid process causing differences of more than 300% in ion uptake efficiency within 1 h, depending on the composition of the pretreatment solution. At concentrations above 0.1 mM the rate of uptake of rubidium in the root was higher than the net potassium uptake necessary for adequate growth. The rate of sulphate uptake was related to potassium uptake but not to phosphate uptake. It is suggested that ion uptake of the roots is regulated by internal factors as well as by direct interactions between the medium and the absorbing surfaces.  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