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 elongation growth of the Avena first internode segments was studied in the presence of one or several of the following growth substances: indoleacetic acid (IAA), 6-fur-furylamino purine (FAP, kinetin), 6-benzylamino purine (BAP), gibberellin A3 (GA3) and A4+7 (GA4+7), and abscisic acid (ABA). The cytokinins at concentrations of 10?7 to 10?6M stimulated growth with 4 to 6 per cent but this effect was not statistically significant. Concentrations higher than 5 × 10?6M inhibited growth. FAP and BAP (from 10?8M to 10?6M) had no significant interaction with any other growth substance used. The two-factor interactions of IAA × ABA, IAA × GA3, and GA3× ABA, as well as the three-factor interaction IAA × ABA × GA3 were significant. However, the IAA × ABA interaction was significant only when high concentration (10?6M) of ABA was used. The growth inhibition produced by 10?7 and 10?6M ABA was overcome by about equimolar concentrations of IAA. The stimulation of growth by GA3 and GA4+7 (10?9 to 10?7M) was prevented by simultaneous application of ABA, and it was reduced significantly by application of IAA (10?7 to 10?8M). GA3 at 10?8M combined with different concentrations of IAA gave slightly higher elongation than IAA alone but the observed values were significantly lower than expected assuming independent additive action.  相似文献   

2.
Kinetin has a stimulating effect in the Avena straight-growth test. The action of different concentrations of kinetin, 2.5 × 10?7, 2.5 × 10?6 and 2.5 × 10?5M, in combination with different concentrations of IAA was studied in this test. It was shown that the effect of low IAA concentrations, 0.25 × 10?7 and 1 × 10?7M, was strongly enhanced by the addition of all the kinetin concentrations investigated. The effect of the highest IAA concentrations, 25 × 10?7 and 100 × 10?7M, on the other hand, was inhibited relatively strongly by the highest employed concentration of kinetin. The results are explained as due to a kinetin-produced increase of auxin in the coleoptile segment, which in combination with low IAA concentrations can lead to a growth stimulation and with high IAA concentrations to a growth inhibition. Since kinetin in purification and chromatography of auxin can partly follow IAA, thereby affecting the quantitative yield, it is emphasized that, prior to the test, auxin extracts containing cytokinins should be freed from the latter by, for example, gel filtration or paper electrophoresis.  相似文献   

3.
The shoot growth and fresh weight of Mentha piperita grown in soil were stimulated at concentrations of 1.26 × 10?5M to 7.77 × 10?4M phosfon (2,4-dichlorobenzyl tributyl phosphonium chloride) while higher concentrations resulted in retardation of growth. Concentrations of 6.30 × 10?7M to 3.78 × 10?5M caused retardation of growth in mineral nutrient solution, and even death at the highest concentrations. However, when the M. piperita plants were grown in mineral nutrient solutions at concentrations of phosfon which had been sequentially lowered from 2.52 × 10?8M to 2.52 × 10?12M, the shoot growth and fresh weight were stimulated as in the case of plants grown in phosfon treated soil.  相似文献   

4.
The effect of cytokinin, kinetin, on abscisic acid (dormin) inhibition of α-amylase synthesis and growth in intact barley seed was investigated. Abscisic acid at 5 × 10?5M nearly completely inhibited growth response and α-amylase synthesis in barley seed. Kinetin reversed to a large extent abscisic acid inhibition of α-aniylase synthesis and coleoptile growth. The response curves of α-amylase synthesis and coleoptile growth in presence of a fixed amount of abscisic acid (6 × l0?6M) and increasing concentrations of kinetin (from 5 × l0?7M to 5 × 10?5 M) showed remarkable similarity. Kinetin and abscisic acid caused synergistic inhibition of root growth. Gibberellic acid was far less effective than kinetin in reversing abscisic acid inhibition of α-amylase synthesis and coleoptile growth. A combination of kinetin and gibberellic acid caused nearly complete reversal of abscisic acid inhibition of α-amylase synthesis but not the abscisic acid inhibition of growth. The results suggest that factors controlling α-amylase synthesis may not have a dominant role in all growth responses of the seed. Kinetin possibly acts by removing the abscisic acid inhibition of enzyme specific sites thereby allowing gibberellic acid to function to produce α-amylase.  相似文献   

5.
Inhibition of bovine erythrocyte acetylcholinesterase (free and immobilized on controlled pore glass) by separate and simultaneous exposure to malathion and malathion transformation products which are generally formed during storage or through natural or photochemical degradation was investigated. Increasing concentrations of malathion, its oxidation product malaoxon, and its isomerisation product isomalathion inhibited free and immobilized AChE in a concentration-dependent manner. KI, the dissociation constant for the initial reversible enzyme inhibitor-complex, and k3, the first order rate constant for the conversion of the reversible complex into the irreversibly inhibited enzyme, were determined from the progressive development of inhibition produced by reaction of native AChE with malathion, malaoxon and isomalathion. KI values of 1.3 × 10? 4 M? 1, 5.6 × 10? 6 M? 1 and 7.2 × 10? 6 M? 1 were obtained for malathion, malaoxon and isomalathion, respectively. The IC50 values for free/immobilized AChE, (3.7 ± 0.2) × 10? 4 M/(1.6 ± 0.1) × 10? 4, (2.4 ± 0.3) × 10? 6/(3.4 ± 0.1) × 10? 6 M and (3.2 ± 0.3) × 10? 6 M/(2.7 ± 0.2) × 10? 6 M, were obtained from the inhibition curves induced by malathion, malaoxon and isomalathion, respectively. However, the products formed due to photoinduced degradation, phosphorodithioic O,O,S-trimethyl ester and O,O-dimethyl thiophosphate, did not noticeably affect enzymatic activity, while diethyl maleate inhibited AChE activity at concentrations > 10 mM. Inhibition of acetylcholinesterase increased with the time of exposure to malathion and its inhibiting by-products within the interval from 0 to 5 minutes. Through simultaneous exposure of the enzyme to malaoxon and isomalathion, an additive effect was achieved for lower concentrations of the inhibitors (in the presence of malaoxon/isomalathion at concentrations 2 × 10? 7 M/2 × 10? 7 M, 2 × 10? 7 M/3 × 10? 7 M and 2 × 10? 7 M/4.5 × 10? 7 M), while an antagonistic effect was obtained for all higher concentrations of inhibitors. The presence of a non-inhibitory degradation product (phosphorodithioic O,O,S-trimethyl ester) did not affect the inhibition efficiencies of the malathion by-products, malaoxon and isomalathion.  相似文献   

6.
The growth response (increase in weight) of cultured explants from seedling date (Phoenix dactylifera L.) and mature coconut (Cocos nucifera L. cv. Malayan Dwarf) palms to source and concentration of organic nitrogen. carbohydrate, auxins, cytokinins and gibberellins was examined. Growth was strongly stimulated by the presence of auxins (10?7 to 10?6M), cytokinins (10?6 to 10?5M), high concentrations of sucrose (0.2 M), and in the absence of NH4Cl, by organic sources of reduced nitrogen. Higher concentrations of auxin (2,4-D or NAA at 10?6 to 10?5M) which still stimulated growth of Phoenix tissue, proved inhibitory to growth of freshly excised Cocos tissues. Explants from both palms initiated roots when subcultured on a medium with increased levels of auxin (NAA, 2.5 × 10?6 to 2.5 × 10?5M) and reduced levels of cytokinin (6-BAP, 5 × 10?8M). Isolated roots excised from these explants continued growth and produced new laterals when subcultured on media with GA3 (5 × 10?7M) and reduced levels of auxin, cytokinin, and either minerals or sucrose.  相似文献   

7.
The growth response of Coelastrum proboscideum Bohlin to cadmiun (3 × 10?9M to 2.4 × 10?7M) was studied. The inorganic media used varied in zinc concentration (1.3 × 10?7M to 3.6 × 10?6M). The data were evaluated by factorial analyses. The influence of zine on the growth depression by cadmium depended on the light conditions (16:8 h light:dark cycles or 24 h continuous illumination periods). Intermittent illumination caused a negative interaction of zinc and cadmium in contrast to a positive interaction or additive effects of these elements during continuous illumination.  相似文献   

8.
Pea plants (Pisum sativum L. cv. Alaska) were grown from seeds for eleven days at different irradiances. Cuttings were then excised and rooted at 16 W × m?2. Gibberellic acid (GA3, 10?11 to 10?3M) was applied to the cuttings immediately after excision. Cuttings from stock plants grown at the highest level of irradiance (38 W × m?2) formed the lowest number of roots. An increasing number of roots per cutting was obtained by decreasing the irradiance to the stock plants. In cuttings from stock plants grown at low irradiances, low concentrations of GA3 (10?8 and 10?7M) promoted root formation further. No effect on rooting by these GA3 concentrations was observed when applied to cuttings originating from stock plants grown at the high irradiances. Root formation in all cuttings was inhibited by GA3 at concentrations higher than 10?6M. The degree of inhibition by GA3 was influenced by the irradiance pretreatment and was increased with an increase in the irradiance during the stock plant growth. Seeds from different years produced cuttings with different response patterns regarding the irradiance and GA3 effects on rooting.  相似文献   

9.
Triiodothyronine, reverse triiodothyronine and thyroxine were found to inhibit 125I labelled thyrotropin binding to human thyroid plasma membranes in vitro. Both the thyrotropin binding and the effect of the above iodoamino-acids on this binding were pH, temperature and time dependent, 50% inhibition of thyrotropin binding was observed at 2×10?7M concentration of reverse triiodothyronine or thyroxine and at 1.1 × 10?6M concentration of triiodothyronine. The kinetic studies of thyrotropin binding revealed that the maximal capacity of receptor sites for the pituitary hormone is unaffected by the presence of thyroid hormones. On the other hand the association and dissociation constants for thyrotropin binding changed when iodoaminoacids were present in the incubation medium /Ka 8.13 × 107M?1 vs 1.6 × 108M?1 and Kd 1.14 × 10?8M vs 4.55 × 10?9M respectively, depending on the pH/. The double reciprocal plots showed competitive mechanism of inhibition. The present study suggest that triiodothyronine, reverse triiodothyronine and thyroxine are able to modify the thyrotropin binding to membrane receptors.  相似文献   

10.
—The hydrolysis of ThTP by rat brain membrane-bound ThTPase is inhibited by nucleoside diphosphates and triphosphates. ATP and ADP are most effective, reducing hydrolysis by 50% at concentrations of 2 × 10?5m and 7·5 × 10?5m respectively. Nucleoside monophosphates and free nuclcosides as well as Pi have no effect on enzyme activity. ThMP and ThDP also fail to inhibit hydrolysis in concentrations up to 5 × 10?3m . Non-hydrolysable methylene phosphate analogs of ATP and ADP were used in further kinetic studies with the ThTPase. The mechanism of inhibition by these analogs is shown to be of mixed non-competitive nature for both compounds. An observed Ki, of 4 × 10?5m for the ATP analog adenosine-PPCP and 9 × 10?5m for the ADP analog adenosine-PCP is calculated at pH 6·5. Formation of the true enzyme substrate, the [Mg2+. ThTP] complex, is not significantly affected by concentrations of analogs producing maximal (>95%) inhibition of enzyme activity. Likewise the relationships between pH and observed Km and pH and Vmax are not shifted by the presence of similar concentrations of inhibitor.  相似文献   

11.
Glucose-6-phosphate dehydrogenase (E.C. 1.1.1.49) was partially purified by fractionation with ammonium sulfate and phosphocellulose chromatography. The Km value for glucose-6-phosphate is 1.6 × 10?4 and 6.3 × 10?4M at low (1.0–6.0 × 10?4M) and high (6.0–30.0 × 10?4M) concentrations of the substrate, respectively. The Km value for NADP+ is 1.4 × 10?5M. The enzyme is inhibited by NADPH, 5-phosphoribosyl-1-pyrophosphate, and ATP, and it is activated by Mg2+, and Mn2+. In the presence of NADPH, the plot of activity vs. NADP+ concentration gave a sigmoidal curve. Inhibition of 5-phosphoribosyl-1-pyrophosphate and ATP is reversed by Mg2+ or a high pH. It is suggested that black gram glucose-6-phosphate dehydrogenase is a regulatory enzyme of the pentose phosphate pathway.  相似文献   

12.
The effect of the insecticide Tanrec® at concentrations of 3.0 × 10?7, 3.0 × 10?2, and 3.0 × 10?1 mg/L (as of imidacloprid) on Daphnia magna Straus has been studied. An acute toxic effect of this insecticide at a concentration of 3.0 × 10?1 mg/L and a depressive effect at concentrations of 3.0 × 10?2 mg/L and 3.0 × 10?7 have been revealed. A damaging effect of Tanrec was revealed during the stage of early development of studied crustaceans. This effect was manifested in the inhibition of the growth of oocytes, abnormal functioning of the intestine, retardation of body growth, and pathological changes in tissues.  相似文献   

13.
In search for the mechanism of insecticidal action of nicotinoids, the kinetics of house fly head cholinesterase inhibition by nicotine were studied to determine the type of inhibition. The pH dependency of inhibition was interpreted in terms of protonation of nitrogen atom in the molecule and the inhibition was shown to be the mixed type closing to competitive type. The Michaelis constants are 3.5 × 10?4 M and 4.1 × 10?4 M, while the apparent inhibition constants obtained are 1.0 × 10?3 M and 2.3 × 10?3 M at pH 7.4 and 8.4, respectively. The type of the inhibition by nicotine monomethiodide carring univalent cation was competitive and the apparent inhibition constant is 1.5 × 10?4 M. These data indicated that the cationic head of nicotinium ion interacts with the anionic site in the active center of cholinesterase.  相似文献   

14.
The elongation and geotropic responses of coleoptile sections of Avena sativa L. to various concentrations of 4-amino-3,5,6-trichloropicolinic acid (Tordon) proved to be qualitatively similar to those previously reported for 2,3,6-trichlorobenzoic acid (TCBA). Tordon stimulated growth in a range of concentrations from 1 × 10?6 to 1 × 10?4M but higher concentrations were inhibitory. Geotropic curvature was extensively depressed by 1 × 10?5 and 1 × 10?4M Tordon, concentrations which accelerated elongation. A similar differential effect has been reported for TCBA and other auxins. Several other picolinic acids and related compounds were tested, but only very slight responses were noted.  相似文献   

15.
Picrotoxin, 1 × 10?5M to 1.6 × 10?3M, had little or no effect on the amplitude of intracellularly recorded excitatory junctional potentials (EJPs) at extracellular calcium concentrations [Ca2+]0 ranging from 0.5 to 15 mM. The slope of the log EJP vs. log[Ca2+]0 relationship was approximately 1 with or without picrotoxin. The reduction of EJP amplitude resulting from the addition of 5 × 10?5M GABA was largely reversed by 10?5M picrotoxin.  相似文献   

16.
Datura innoxia Mill. callus cultures formed shoots in 2–4 weeks on media containing; a) gibberellic acid, b) indoleacetic acid, c) low concentrations of naphthylacetic acid, d) low concentrations of 2,4-dichlorophenoxyacetic acid, e) benzylaminopurine, f) no growth substance. Benzylaminopurine promoted shoot differentiation. Gibberellic acid inhibited shoot formation weakly, but inhibited proper leaf blade formation. Root differentiation was rare. The callus cultures of Datura innoxia grew rapidly (100-fold in 4 weeks) on a slightly modified Murashige and Skoog medium (0.5 mg/l thiamin · HCl, pH 5.5, no glycine) in light at 30°C. Callus grew well on any single one of the growth substances NAA (10?5M), 2,4-D (10?6M) or BAP (3 × 10?6M). Growth was less and more erratic on GA or IAA. The callus cultures did not grow significantly better when BAP was combined with one of the auxins or with GA.  相似文献   

17.
《Free radical research》2013,47(9):1150-1156
Abstract

Oxidation of tyrosine moieties by radicals involved in lipid peroxidation is of current interest; while a rate constant has been reported for reaction of lipid peroxyl radicals with a tyrosine model, little is known about the reaction between tyrosine and alkoxyl radicals (also intermediates in the lipid peroxidation chain reaction). In this study, the reaction between a model alkoxyl radical, the tert-butoxyl radical and tyrosine was followed using steady-state and pulse radiolysis. Acetone, a product of the β-fragmentation of the tert-butoxyl radical, was measured; the yield was reduced by the presence of tyrosine in a concentration- and pH-dependent manner. From these data, a rate constant for the reaction between tert-butoxyl and tyrosine was estimated as 6?±?1 × 107 M?1 s?1 at pH 10. Tyrosine phenoxyl radicals were also monitored directly by kinetic spectrophotometry following generation of tert-butoxyl radicals by pulse radiolysis of solutions containing tyrosine. From the yield of tyrosyl radicals (measured before they decayed) as a function of tyrosine concentration, a rate constant for the reaction between tert-butoxyl and tyrosine was estimated as 7?±?3 × 107 M?1 s?1 at pH 10 (the reaction was not observable at pH 7). We conclude that reaction involves oxidation of tyrosine phenolate rather than undissociated phenol; since the pKa of phenolic hydroxyl dissociation in tyrosine is ~ 10.3, this infers a much lower rate constant, about 3 × 105 M?1 s?1, for the reaction between this alkoxyl radical and tyrosine at pH 7.4.  相似文献   

18.
The effects of nitrogen (N) nutrition on growth, N uptake and leaf osmotic potential of rice plants (Oryza sativa L. ev. IR 36) during simulated water stress were determined. Twenty-one-day-old seedlings in high (28.6 × 10 ?4M) and low (7.14 × 10 4M) N levels were exposed to decreased nutrient solution water potentials by addition of polyethylene glycol 6000. The roots were separated from the solution by a semi-permeable membrane. Nutrient solution water potential was ?0.6 × 105 Pa and was lowered stepwise to ?1 × 105, ?2 × 105, ?4 × 105 and ?6 × 105 Pa at 2-day intervals. Plant height, leaf area and shoot dry weight of high and low nitrogen plants were reduced by lower osmotic potentials of the root medium. Osmotic stress caused greater shoot growth reduction in high N than in low N plants. Stressed and unstressed plants in 7.14 × 104M N had more root dry matter than the corresponding plants in 28.6 × 104M N. Dawn leaf water potential of stressed plants was 1 × 105 to 5.5 × 105 Pa lower than nutrient solution water potential. Nitrogen-deficient water-stressed plants, however, maintained higher dawn leaf water potential than high nitrogen water-stressed plants. It is suggested that this was due to higher root-to-shoot ratios of N deficient plants. The osmotic potentials of leaves at full turgor for control plants were about 1.3 × 105 Pa higher in 7.14 × 10?4M than in 28.6 × 10?4M N and osmotic adjustment of 2.6 × 105 and 4.3 × 105 Pa was obtained in low and high N plants, respectively. The nitrogen status of plants, therefore, affected the ability of the rice plant to adjust osmotically during water stress. Plant water stress decreased transpiration and total N content in shoots of both N treatments. Reduced shoot growth as a result of water stress caused the decrease in amount of water transpired. Transpiration and N uptake were significantly correlated. Our results show that nitrogen content is reduced in water-stressed plants by the integrated effects of plant water stress per se on accumulation of dry matter and transpiring leaf area as well as the often cited changes in soil physical properties of a drying root medium.  相似文献   

19.
T. viride ITCC 1433 synthesizes a two component system for the hydrolysis of cellobiose and cellooligodextrins. 80% of the total activity are solubilized during growth. The large protein (A), mol. weight 98 000 d, is glycosylated and slightly acidic (pH = 6.1). The smaller protein (B), mol. weight 70 000 d, is unglycosylated and neutral (pH = 7.2). Both proteins form a two-step system where β-glucosidase A is active at low substrate concentrations (KM = 2.3 × 10?4 M cellobiose) while β-glucosidase B covers the range of 10-fold higher cellobiose concentrations (KM = 1.8 × 10?3 M). The enzymes are fairly stable with a residual activity of 70% at 50°C after 24 h.  相似文献   

20.
Abstract: In this study, we have investigated the effect of mivazerol, [3-(1H-imidazol-4-yl)methyl-1]-2-hydroxy-benzamide hydrochloride, a new α2-agonist lacking hypotensive properties and a potential anti-ischemic drug, on the evoked release of norepinephrine, aspartate, and glutamate in tissue preparations from hippocampus, spinal cord T1–T5 section, rostrolateral ventricular medulla, and nucleus tractus solitarii of the brainstem of rat. A simple and efficient in vitro procedure to study pharmacologically the release of norepinephrine and glutamate is described. Tissues were chopped into (0.3 × 0.2 × 0.2 mm3) sections and the resulting minces were used for this study. Exposure to KCl (10–75 mM) for 5 min served as a stimulus for the release response. One, S (for aspartate and for glutamate release), or two such stimuli, S1 and S2 (for norepinephrine release) were conducted. The release of norepinephrine (+150% above baseline) was inhibited in a dose-dependent manner by mivazerol in hippocampus (IC50 = 1.5 × 10?8M), spinal cord (IC50 = 5 × 10?8M), rostrolateral ventricular medulla (IC50 = 10?7M), and nucleus tractus solitarii (IC50 = 7.5 × 10?8M), and by clonidine in hippocampus (IC50 = 5 × 10?8M), spinal cord (IC50 = 4.5 × 10?8M), rostrolateral ventricular medulla (IC50 = 2.5 × 10?7M), and nucleus tractus solitarii (IC50 = 10?7M). This effect was counteracted by the selective α2-antagonists yohimbine and rauwolscine. A significant glutamate and aspartate release response was also induced by KCl (35 mmol/L) in hippocampus (+250 and +135%, respectively) and spinal cord (+120 and +55%, respectively), in vitro. However, neither mivazerol nor clonidine, at doses up to 10 µM, had any significant effect on KCl-induced glutamate release in spinal cord, whereas mivazerol blocked completely the release of both amino acids in hippocampus and only the release of aspartate in spinal cord. On the other hand, clonidine (1 µM) was only effective in reducing by 40% the release of aspartate in hippocampus. These data indicate that (1) inhibition of KCl-induced norepinephrine release by mivazerol is mediated by its action on α2-adrenergic receptors; (2) at concentrations selective for α2-adrenergic receptors, only mivazerol was effective in blocking the KCl-induced glutamate release in hippocampal tissue; and (3) at the same concentrations, both mivazerol and clonidine were unable to inhibit glutamate release in the spinal cord. These data suggest that prevention of hyperadrenergic activity by mivazerol in perioperative patients may be mediated through its effect on the release of norepinephrine and/or the release of glutamate and aspartate in regions of the CNS that are involved in the control of cardiovascular homeostasis.  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