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 46 毫秒
1.
ABSTRACT The relative capacity of Na+, K+ and Cl- to stimulate germination of spores of the microsporidian Nosema algerae, a pathogen of mosquitoes, was examined by ion substitution experiments. Sodium at 0.1 M was ineffective to produce the high percentage of germination that typically occurs with 0.1 M NaCl (the normal stimulation solution) if Cl- was substituted with the usually impermeant anions SO42-, HPO42-, or the organic acids oxalate, cacodylate, EGTA, MES and HEPES. However, substantial concentration- and pH-dependent germination was seen with Na2SO4 in the 0.2-0.8 M Na+ range. Similar results were obtained with solutions of K+ accompanied by impermeant anions. In contrast, the chloride salts of usually impermeant cations, like choline and triethanolamine, failed to germinate spores even at 0.8 M unless Na+ or K+ was independently added. The presence of 0.5 M choline chloride in the medium reduced the levels of Na2SO4 required to produce germination down to equivalence with those of Na+ in the normal stimulation solution. Monensin, a Na+ ionophore, facilitated the germination induced by a medium-level stimulus (0.04 M NaCl) in sonicated samples. These findings indicate that N. algerae spores germinate in response to the alkali metal cations, while CI- plays a passive role by diffusing to maintain internal electroneutrality during cation influx. A possible mechanism of cation action in spore germination is suggested on the basis of these results and observations on other systems of intracellular motility.  相似文献   

2.
Interactive effects of Fe2(SO4)3 and Na2EDTA on callus induction and plant regeneration in indica rice (Oryza sativa cv Pusa Basmati-1) was investigated. Callus induction and subsequent plant regeneration from seed was obtained on MS medium supplemented with 11.31 µM 2,4-D and 2.68 µM NAA + 8.87 µM BAP respectively. Both the callus induction and the plant regeneration media were supplemented with different levels of Fe2(SO4)3, Na2EDTA and their combinations. Callus induction, its morphogenic potential, average number of regenerated plants as well as the appearance of the regenerants was influenced by the levels of Fe2(SO4)3 and Na2EDTA. Embryogenic callus could be induced in the presence of Fe2(SO4)3 / Na2EDTA. Iron was essential for plant regeneration. Non-chelated iron could induce callus formation as well as plant regeneration, yet the chelated form was more suitable. A higher level of Na2EDTA in the regeneration medium was detrimental. Differential requirement of Fe2(SO4)3 at induction and regeneration level was observed. Method of medium preparation affected the regeneration response. Nearly three fold increase in the number of regenerated plants was achieved with 0.05 mM Fe2(SO4)3 + 0.1 mM Na2EDTA at callus induction and 0.1 mM Fe2(SO4)3+ 0.1 mM Na2EDTA at plant regeneration and standard method of autoclaving.  相似文献   

3.
Sorghum bicolor L. Moench, RS 610, was grown in liquid media salinized with NaCl, KCl, Na2SO4, K2SO4 or with variable mixtures of either NaCl/KCl or Na2SO4/K2SO4 at osmotic potentials ranging from 0 to -0.8 MPa. The purpose was to study the effects of different types and degrees of salinity in growth media on growth and solute accumulation. In 14-day-old plants the severity of leaf growth inhibition at any one level of osmotic potential in the medium increased according to the following order: NaCl < Na2SO4 < KCl = K2SO4. Inhibition of growth by mixtures of Na+ and K+ salts was the same as by K+ salts alone. Roots responded differently. Root growth was not affected by Na+ salts in the range of 0 to -0.2 MPa while it was stimulated by K+ salts. The major cation of leaves was K+ because S. bicolor is a Na+-excluder, while Na+ was the major cation in roots except at low Na+/K+ ratios in media. Anions increased in tissues linearly in relation to total monovalent cation, but not with a constant anion/cation ratio. This ratio increased as the cation concentrations in tissues increased. Sucrose in leaf tissue increased 75 fold in Chloride-plants (plants growing in media in which the only anion of the salinizing salts was Cl?) and 50 fold in Sulphate-plants (the only anion of the salinizing salts was SO42-). Proline increased 60 and 18 fold in Chloride- and Sulphate-plants, respectively, as growth media potentials decreased from 0 to -0.8 MPa. The concentrations of both sucrose and proline were directly proportional to the amount of total monovalent cation in the tissue. Sucrose concentrations began increasing when total monovalent cations exceeded 100 μmol (g fresh weight)?1 (the monovalent cation level in non-stressed plants), but proline did not start accumulating until monovalent cation concentrations exceeded 200 μmol (g fresh weight)?1. Therefore, sucrose seemed to be the solute used for osmotic adjustment under mild conditions of saline stress while proline was involved in osmotic adjustment under more severe conditions of stress. Concentrations of inorganic phosphate, glucose, fructose, total amino acids and malic acid fluctuated in both roots and leaves in patterns that could be somewhat correlated with saline stress and, sometimes, with particular salts in growth media. However, the changes measured were too small (at most a 2–3 fold increase) to be of importance in osmotic adjustment.  相似文献   

4.
Changes in the surface potential, the electrical potential difference between the membrane surface and the bulk aqueous phase were measured with the carotenoid spectral shift which indicates the change of electrical field in the membrane. Chromatophores were prepared from a non-sulfur purple bacterium, Rhodopseudomonas sphaeroides, in a low-salt buffer. Surface potential was changed by addition of salt or by pH jump as predicted by the Gouy-Chapman diffuse double layer theory.When a salt was added at neutral pH, the shift of carotenoid spectrum to shorter wavelength, corresponding to an increase in electrical potential at the outside surface, was observed. The salts of divalent cations (MgSO4, MgCl2, CaCl2) were effective at concentrations lower than those of monovalent cation salts (NaCl, KCl, Na2SO4) by a factor of about 50. Among the salts of monoor divalent cation used, little ionic species-dependent difference was observed in the low-concentration range except that due to the valence of cations. The pH dependence of the salt-induced carotenoid change was explained in terms of the change in surface charge density, which was about 0 at pH 5–5.5 and had negative values at higher pH values. The dependence of the pH jump-induced absorbance change on the salt concentration was also consistent with the change in the charge density. The surface potential change by the salt addition, which was calibrated by H+ diffusion potential, was about 90 mV at the maximum. From the difference between the effective concentrations with salts of mono- and divalent cations at pH 7.8, the surface charge density of (?1.9 ± 0.5) · 10?3 elementary charge per Å2, and the surface potential of about ?100 mV in the presence of about 0.1 mM divalent cation or 5 mM monovalent cation were calculated.  相似文献   

5.
The effects of NaCl and Na2SO4 on photosynthetic pigments, malondialdehyde (MDA), Rubisco activity and superoxide dismutase (SOD) activity were investigated in Kalidium foliatum (Pall.) Moq., which is distributed in the saline soil of Hetao irrigation area in Inner Mongolia China. The K. foliatum plants were treated with NaCl (0, 100, 250, 400 and 500 mM), Na2SO4 (0, 100, 250, 400 and 500 mM) and NaCl + Na2SO4 (1: 1, v/v) (0, 100, 250, 400 and 500 mM of Na+ concentration, 0, 50, 125, 200 and 250 mM of Cl and SO 4 2– concentration) for 10 days. Content of chlorophylls and carotenoids were significantly higher than control at increasing NaCl and Na2SO4 concentration, in contrast, were significantly reduced by higher concentration of NaCl + Na2SO4. Rubisco activity reduced steadily at 100 and 250 mM NaCl, while increased at 400 and 500 mM NaCl. Rubisco activity was significantly higher than control at 100 mM Na2SO4, and was no more change under NaCl + Na2SO4 treatment. The SOD activity increased with increasing NaCl and Na2SO4, and increased at moderate NaCl + Na2SO4 treatment. MDA content was lower than control at 250 mM salt concentration. On the basis of the data obtained, K. foliatum showed resistance to salt such as Na+, Cland SO 4 2– , Rubisco activity in K. foliatum might be more sensitive to salt.  相似文献   

6.
Potassium toxicity to survival and growth of Microcystis has been investigated for the first time by taking photosynthetic parameters and change in internal pH of Microcystis. The concentration of potassium reducing 50% population of Microcystis was found to be 6 mM. At this concentration, the internal pH of cells increased from 7.2 to 9.8 in comparison to control. 6.0 mM concentration of potassium reduced protein content by 44% and generated Na+ efflux of 55% as compared to control. O2 evolution, ATP content and CO2 fixation were found to be very sensitive to above K+ concentration and registered a respective decline of 38, 32 and 36%. PS II was the primary site of action depicting about 35% inhibition at above K+ concentration. PS I and whole electron transport chain were also inhibited but the extent was less pronounced in comparison to PS II. A definite correlation between requirement of Na+ for growth and maintenance of cytoplasmic pH was observed. K+-induced loss of Na+ from cells of Microcystis could result in increase in internal pH, which in turn affects survival, growth, and other physiological parameters of Microcystis. Thus, K+ appears to hold excellent potential for the control of Microcystis blooms in fresh water ponds and lakes.  相似文献   

7.
H P Hopkins  W D Wilson 《Biopolymers》1987,26(8):1347-1355
Enthalpy changes (ΔHB) for the binding of ethidium (a monocation) and propidium (a dication) to calf thymus DNA have been determined calorimetrically in piperazine-N, N′-bis(2-ethanesulfonic acid) buffer with the fluoride ion as the counterion. Heats of dilution for the fluoride salts of ethidium and propidium were substantially less than the corresponding values found for other halide salts of these cations. At a Na+ ion concentrations of 0.019, ΔHB = ?8.3 and ?7.9 ± 0.3 kcal mol?1 for ethidium and propidium, respectively. For these two cations, just as was observed for the naphthalene monoimide (monocation) and diimide (dication) [H. P. Hopkins, K. A. Stevenson, and W. D. Wilson, (1986) J. Sol. Chem. 15 , 563–579], ΔHB is within the same experimental error for both cations. Apparently, charge–charge interactions in DNA–cation complexes produce only small changes in the enthalpy for the system. In the concentration range 0.019–0.207, the ΔHB values for propidium did not depend appreciably on the Na+ ion concentration, and a similar pattern was shown to exist for ethidium. When these results were combined with ΔGB values for the binding of these cations to DNA, we found the variation of ΔSB with Na+ ion concentration to be remarkably close to the predictions of modern polyelectrolyte theory, i.e., propidium binding to DNA causes approximately twice as many Na+ ions to be released into the bulk solution as does the binding of ethidium. The much stronger binding of propidium, relative to ethidium, at low ionic strengths is thus seen to be primarily due to entropic effects.  相似文献   

8.
The permeation of K+ and Na+ through the pore of a K+ channel from the plasma membrane of rye roots was studied in planar 1-palmitoyl-2-oleoyl phosphatidylethanolamine bilayers. The pore contains at least two ion-binding sites which can be occupied simultaneously. This was indicated by: (i) biphasic relationships with increasing cation concentration of both channel conductance at the zero-current (reversal) potential of the channel (E rev) and unitary-current at a specified voltage and (ii) a decline in E rev in the presence of equimolar Na+ (cis):K+ (trans) as the cation concentration was increased. To determine the spatial characteristics and energy profiles for K+ and Na+ permeation, unitary-current/ voltage data for the channel were fitted to a three energy-barrier, two ion-binding site (3B2S) model. The model allowed for simultaneous occupancy of binding sites and ionic repulsion within the pore, as well as surface potential effects. Results suggested that energy peaks and energy wells (ion binding sites) were situated asymmetrically within the electrical distance of the pore, the trans energy-well being closer to the center of the pore than its cis counterpart; that the energy profile for K+ permeation differed significantly from that of Na+ in having a higher cis energy peak and a deeper cis energy well; that cations repelled each other within the pore and that vestibule surface charge was negligible. The model successfully simulated various aspects of K+ and Na+ permeation including: (i) the complexities in current rectification over a wide range of contrasting ionic conditions; (ii) the biphasic relationships with increasing cation concentration of both channel conductance at E rev and unitary-current at a specified voltage; (iii) the decline in E rev in equimolar Na+ (cis):K+ (trans) as cation concentrations were increased and (iv) the complex relationships between mole fraction and E rev at total cation concentrations of 100 and 300 mm.We thank Prof. O. Alvarez (Universidad de Chile, Santiago, Chile) for supplying the computer program AJUSTE and Prof. D. Sanders (University of York, UK), Prof. D. Gradmann and Dr. G. Thiel (University of Göttingen, Germany) for stimulating ideas. This work was supported by the Agricultural and Food Research Council.  相似文献   

9.
1. To reveal the role of aquatic heterotrophic bacteria in the process of development of Microcystis blooms in natural waters, we cocultured unicellular Microcystis aeruginosa with a natural Microcystis‐associated heterotrophic bacterial community. 2. Unicellular M. aeruginosa at different initial cell densities aggregated into colonies in the presence of heterotrophic bacteria, while axenic Microcystis continued to grow as single cells. The specific growth rate, the chl a content, the maximum electron transport rate (ETRmax) and the synthesis and secretion of extracellular polysaccharide (EPS) were higher in non‐axenic M. aeruginosa than in axenic M. aeruginosa after cell aggregation, whereas axenic and non‐axenic M. aeruginosa displayed the same physiological characteristic before aggregation. 3. Heterotrophic bacterial community composition was analysed by PCR–denaturing gradient gel electrophoresis (PCR–DGGE) fingerprinting. The biomass of heterotrophic bacteria strongly increased in the coinoculated cultures, but the DGGE banding patterns in coinoculated cultures were distinctly dissimilar to those in control cultures with only heterotrophic bacteria. Sequencing of DGGE bands suggested that Porphyrobacter, Flavobacteriaceae and one uncultured bacterium could be specialist bacteria responsible for the aggregation of M. aeruginosa. 4. The production of EPS in non‐axenic M. aeruginosa created microenvironments that probably served to link both cyanobacterial cells and their associated bacterial cells into mutually beneficial colonies. Microcystis colony formation facilitates the maintenance of high biomass for a long time, and the growth of heterotrophic bacteria was enhanced by EPS secretion from M. aeruginosa. 5. The results from our study suggest that natural heterotrophic bacterial communities have a role in the development of Microcystis blooms in natural waters. The mechanisms behind the changes of the bacterial community and interaction between cyanobacteria and heterotrophic bacteria need further investigations.  相似文献   

10.
L G Foe  J L Trujillo 《Life sciences》1979,25(17):1529-1538
The monovalent cations NH4+, K+, and Rb+ activate pig liver phosphofructokinase by increasing the maximal velocity. In the presence of these cations the enzyme retains sigmoid kinetics with respect to fructose-6-phosphate. However, these cations bring about a decrease in the [S]0.5 for fructose-6-phosphate to an extent directly proportional to their ionic volumes. The apparent dissociation constants of NH4+, K+, and Rb+ for the enzyme at 0.5 mM ATP and 4 mM Fru6P are 0.2 mM, 8 mM, and 15 mM, respectively. The maximal velocity of the enzyme in the presence of saturating concentrations of Rb+ is about 70% of that seen with NH4+ or K+. The monovalent cations Li+, Na+, and Cs+ inhibit the enzyme at high concentrations (> 50 mM) by decreasing the maximal velocity. Although the efficiency of inhibition by these cations qualitatively increases with decreasing size, there is no obvious quantitative relationship between efficiency of inhibition and any parameter of ionic size.  相似文献   

11.
Summary The kinetics of K+ and Na+ transport across the membrane of large unilamellar vesicles (L.U.V.) were compared at two pH's, with two carriers: (222)C 10-cryptand (diaza-1, 10-decyl-5-hexaoxa-4,7,13,16,21,24-bicyclo[8.8.8.]hexacosane) and valinomcyin, i.e. an ionizable macrobicyclic amino polyether and a neutral macrocyclic antibiotic. The rate of cation transport by (222)C10 saturated as cation and carrier concentrations rose. The apparent affinity of (222)C10 for K+ was higher and less pH dependent than that for Na+ but resembled the affinity of valinomycin for K+. The efficiency of (222)C10 transport of K+ decreased as the pH fell and the carrier concentration rose, and was about ten times lower than that of valinomycin. Noncompetitive K+/Na+ transport selectivity of (222)C10 decreased as pH, and cation and carrier concentrations rose, and was lower than that of valinomycin. Transport of alkali cations by (222)C10 and valinomycin was noncooperative. Reaction orders in cationn(S) and carrierm(M) varied with the type of cation and carrier and were almost independent of pH;n(S) andm(M) were not respectively dependent on carrier or cation concentrations. The apparent estimated constants for cation translocation by (222)C10 were higher in the presence of Na+ than of K+ due to higher carrier saturation by K+, and decreased as pH and carrier concentration increased. Equilibrium potential was independent of the nature of carrier and transported cation. Results are discussed in terms of the structural, physicochemical and electrical characteristics of carriers and complexes.  相似文献   

12.
Dissociation of the (Na++K+)-ATPase ouabain complex, formed presence of Mg2+ and inorganic phosphate (Complex II), is inhibited by Mg2+ (21–45%) and the alkali cations Na+ (25–59%) and K+ (27–75%) when kidney cortex tissue (bovine, rabbit, guinea pig) is the enzyme source. Choline chloride at 200 mM, equivalent to the highest concentration of NaCl tested, does not inhibit. Dissociation of Complex II from brain cortex (bovine, rat, rabbit) or heart muscle (rabbit) is much less inhibited: 0–11% by Na+ and 11–19% by K+. The degree of inhibition is not directly related to the size of the dissociation rate constant (k?) of the various complexes, but rather to the extent of interaction between the cation and ouabain binding sites for these tissues.Inhibition curves for Na+ and K+ are sigmoidal. Half-maximal inhibition for rabbit brain and kidney cortex is at 30–40 mM Na+ and 6–10 mM K+, and the maximally inhibitory concentrations are 50–150 and 15–20 mM, respectively. Maximal inhibition by Na+ or K+ for these tissues is the same. For guinea pig kidney cortex Na+ and K+ are almost equally effective, but 150 mM K+ or 200 mM Na+ are still not saturating, and inhibition curves indicate high- and low-affinity binding sites for the alkali cations.The inhibition curve for Mg2+ is not sigmoidal. In the kidney preparations Mg2+ inhibits half-maximally at 0.4-0.5 mM, maximally at 1–3 mM. Maximal inhibition by Mg2+ is higher than by Na+ or K+ for rabbit cortex and lower for guinea pig kidney cortex.There is no competition or additivity among the cations, indicating the existence of different binding sites for Mg2+ and the alkali cations.Complex II differs in stability, in the extent of inhibition, in the dependence of inhibition on the cation concentration and in the absence of antagonism between Na+ and K+, from the ouabain complex formed via phosphorylation by ATP (Complex I). This indicates that the phosphorylation states for the complexes are clearly different.  相似文献   

13.
Gerke  I.  Zierold  K.  Weber  J.  Tardent  P. 《Hydrobiologia》1991,216(1):661-669
The spatial distribution of cations was assayed qualitatively and quantitatively in tentacular nematocytes of Hydra vulgaris in a scanning transmission electron microscope by means of x-ray microanalysis performed on 100 nm thick freeze-dried cryosections. The matrix of undischarged cysts (stenoteles, desmonemes and isorhizas) was found to contain mainly K+. In isolated nematocysts of Hydra the intracapsular potassium can be readily substituted by practically any other mono- and divalent cation (Na+, NH4 +, Mn2+, Co2+, Mg2+, Ca2+, Fe2+) all, except Fe2+, without impairing the ability of the cyst to respond to the chemical triggering with dithioerythritol or proteases. Monovalent cations increase the osmotically generated intracapsular pressure compared to divalent ions.  相似文献   

14.
A recombinant hybrid of manganese dependent-superoxide dismutase of Staphylococcus equorum and S. saprophyticus has successfully been overexpressed in Escherichia coli BL21(DE3), purified, and characterized. The recombinant enzyme suffered from degradation and aggregation upon storage at ?20 °C, but not at room temperature nor in cold. Chromatographic analysis in a size exclusion column suggested the occurrence of dimeric form, which has been reported to contribute in maintaining the stability of the enzyme. Effect of monovalent (Na+, K+), divalent (Ca2+, Mg2+), multivalent (Mn2+/4+, Zn2+/4+) cations and anions (Cl?, SO4 2?) to the enzyme stability or dimeric state depended on type of cation or anion, its concentration, and pH. However, tremendous effect was observed with 50 mM ZnSO4, in which thermostability of both the dimer and monomer was increased. Similar situation was not observed with MnSO4, and its presence was detrimental at 200 mM. Finally, chelating agent appeared to destabilize the dimer around neutral pH and dissociate it at basic pH. The monomer remained stable upon addition of ethylene diamine tetraacetic acid. Here we reported unique characteristics and stability of manganese dependent-superoxide dismutase from S. equorum/saprophyticus.  相似文献   

15.
A protein complex (PC) suspension exhibits asymmetric biooxidation activities in the absence of any added cofactor such as NAD(P)+ or FAD. It can be extracted from pea protein (PP)‐gel (PP encapsulated with Ca2+ alginate gel and aerated in air for several hours) using hot water by rotary shaking and powdered by the following three steps: (1) forming precipitates from the suspension using 30% (w/v) aqueous (NH4)2SO4, (2) crosslinking the precipitates with 0.25% (v/v) GA, and (3) preparing the cross‐linked powder by freeze‐drying. The cross‐linked PC (CLPC) performed asymmetric oxidation of the toward (R)‐isomers of rac‐ 1 and rac ‐2 in 50 mM glycine–NaOH (pH 9.0) buffer/DMSO cosolvent [2.07% (v/v)] with high enantioselectivity; thus, the (S)‐isomers can be obtained in greater than 99% ee from the corresponding racp‐substituted naphthyl methyl carbinol (rac‐ 1 and rac ‐2 ). The CLPC activity was not only competitively inhibited by addition of either 1.0 mM ZnCl2 or a chelating agent such as 1.0 mM EDTA but also denatured by pretreatments: autoclaving at 121°C (20 min) or using 6.0 M guanidine–HCl containing 50 mM DTT. These results indicated that the PC catalytic process may utilize an electron transfer system incorporating a redox cation (e.g., Fe2+ ? Fe3+ or Zn). Therefore, the newly introduced CLPC can asymmetrically oxidize the substrates without the addition of any cofactor resulting in a low‐cost organic method. Overall, our results show that the CLPC is an easily prepared, low‐cost reagent that can function under mild conditions and afford stereoselectivity, regioselectivity, and substrate specificity. © 2012 American Institute of Chemical Engineers Biotechnol. Prog., 28: 953–961, 2012  相似文献   

16.
Abstract. Leaching of inorganic cations (K+, Mg2+) and in some cases of inorganic anions and sugars from detached twigs and single needles of spruce Picea abies L. Karst.) in the presence of acid rain (H2SO4, 1 mol m?3) or salt solutions (Na2 SO4, 1 mol m?3) was examined under laboratory conditions. Cation leaching (as percentage of the total water soluble ion content of the tissue per hour) was: K+: 0.01-0.02%; Mg2+: 0.005-0.01%; Ca2+: 0.1-0.2%. Leaching rates of anions were even lower than that and concentrations in the leachate were often below the detection limit of anion chromatography. Spraying with H2SO4 (pH 2.95, 1 mol m?3) increased leaching only transiently. Similar effects were found when Na2SO4 was used instead of H2SO4. The transiently enhanced leaching was apparently due to H+/cation or cation/cation exchange at the twig or leaf surfaces. Feeding of K+ or Al3+ through the stems increased leaching of all cations within a few hours, again demonstrating rapid ion exchange in the apoplast. Leaching of potassium and magnesium from single needles occurred at similar relative rates as from twigs. Loss of Ca2+ ions, however, was even smaller from needles than from twigs. Apparently, a large part of the Ca2+ lost from twigs originated from the bark and not from the needles. Efflux of ions from longitudinal needle sections was about 1000 times taster than the rates obtained with intact needles, indicating that the cuticle was the main barrier Preventing solute loss. In relation to the total amount of mineral nutrients in trees, leaching is considered to be too small to be the primary cause of damage to trees stressed by acid rain, as has been suggested in the literature.  相似文献   

17.
We studied the peculiarities of permeability with respect to the main extracellular cations, Na+ and Ca2+, of cloned low-threshold calcium channels (LTCCs) of three subtypes, Cav3.1 (α1G), Cav3.2 (α 1H), and Cav3.3 (α1I), functionally expressed in Xenopus oocytes. In a calcium-free solution containing 100 mM Na+ and 5 mM calcium-chelating EGTA buffer (to eliminate residual concentrations of Ca2+) we observed considerable integral currents possessing the kinetics of inactivation typical of LTCCs and characterized by reversion potentials of −10 ± 1, −12 ± 1, and −18 ± 2 mV, respectively, for Cav3.1, Cav3.2, and Cav3.3 channels. The presence of Ca2+ in the extracellular solution exerted an ambiguous effect on the examined currents. On the one hand, Ca2+ effectively blocked the current of monovalent cations through cloned LTCCs (K d = 2, 10, and 18 μM for currents through channels Cav3.1, Cav3.2, and Cav3.3, respectively). On the other hand, at the concentration of 1 to 100 mM, Ca2+ itself functioned as a carrier of the inward current. Despite the fact that the calcium current reached the level of saturation in the presence of 5 mM Ca2+ in the external solution, extracellular Na+ influenced the permeability of these channels even in the presence of 10 mM Ca2+. The Cav3.3 channels were more permeable with respect to Na+ (P Ca/P Na ∼ 21) than Cav3.1 and Cav3.2 (P Ca/P Na ∼ 66). As a whole, our data indicate that cloned LTCCs form multi-ion Ca2+-selective pores, as these ions possess a high affinity for certain binding sites. Monovalent cations present together with Ca2+ in the external solution modulate the calcium permeability of these channels. Among the above-mentioned subtypes, Cav3.3 channels show the minimum selectivity with respect to Ca2+ and are most permeable for monovalent cations. Neirofiziologiya/Neurophysiology, Vol. 38, No. 3, pp. 183–192, May–June, 2006.  相似文献   

18.
All of the common cytochalasins activate superoxide anion release and exocytosis of β-N-acetylglucosaminidase and lysozyme from guinea-pig polymorphonuclear leukocytes (neutrophils) incubated in a buffered sucrose medium. Half-maximal activation of both processes is produced by approx. 2 μM cytochalasin A, C >μM cytochalasin B ? 4–5 μM cytochalasin D, E. While maximal rates of O2? release and extents of exocytosis require extracellular calcium (1–2 mM), replacing sucrose with monovalent cation chlorides is inhibitory to neutrophil activation by cytochalasins. Na+, K+ or choline inhibited either cytochalasin B- or E-stimulated O2? production with IC50 values of 5–10 mM and inhibition occurs whether Cl?, NO3? or SCN? is the anion added with Na+ or K+. Release of β-N-acetylglucosaminidase in control or cytochalasin B-stimulated cells is inhibited by NaCl (IC50 ≈ 10 mM), while cytochalasin E-stimulated exocytosis is reduced less and K+ or choline chloride are ineffective in inhibiting either cytochalasin B- or E-stimulated exocytosis. Release of β-glucuronidase, myeloperoxidase or acid phosphatase from neutrophils incubated in buffered sucrose is not stimulated by cytochalasin B. Stimulation of either O2? or β-N-acetylglucosaminidase release by low concentrations of cytochalasin A is followed by inhibition of each at higher concentrations. It appears that all cytochalasins can activate both NAD(P)H oxidase and selective degranulation of neutrophils incubated in salt-restricted media and that differential inhibition of these two processes by monovalent cations and/or anions is produced at some step(s) subsequent to cytochalasin interaction with the cell.  相似文献   

19.
The addition of fresh serum-containing growth medium to L1210 mouse leukemic cells in culture resulted in a 5-fold increase in ornithine decarboxylase (l-ornithine carboxy-lyase, EC 4.1.1.17) activity. The presence of microtubule disrupting agents (colchine, vinblastine) or cations (5–10 mM K+, Na+ or Mg2+) abolishes this increase of ornithine decarboxylase activity (Chen, K.Y., Heller, J.S. and Canellakis, E.S. (1976) Biochem. Biophys. Res. Commun. 70, 212–219). Based on these observations we proposed that fluctuation in cellular cation concentrations may act as a link between the membrane structure and ornithine decarboxylase. To test this proposal, we studied the effects of selective membrane perturbing agents such as ionophores and local anesthetics, on the serum-stimulated increase of ornithine decarboxylase activity in L1210 cells. Among the six inonophores tested, valinomycin was the most potent one, with I50 value (concentration that gives 50% inhibition of orthinine decarbocylase activity) of 6·10?9 M. Dibucaine and tetracaine were also effective inhibitors at 10?4?10?5 M. The I50 values of valinomycin on the protein synthesis and RNA synthesis, however, were greater than 1·10?6 M. These results substantiate the notion that ornithine decarboxylase activity can be regulated at plasma membrane level and such regulation is related to the perturbation of cellular cation pools.  相似文献   

20.
Kochia sieversiana (Pall.) C. A. M., a naturally alkali-resistant halophyte, was chosen as the test organism for our research. The seedlings of K. sieversiana were treated with varying (0–400 mM) salt stress (1:1 molar ratio of NaCl to Na2SO4) and alkali stress (1:1 molar ratio of NaHCO3 to Na2CO3). The concentrations of various solutes in fresh shoots, including Na+, K+, Ca2+, Mg2+, Cl, SO42−, NO3, H2PO3, betaine, proline, soluble sugar (SS), and organic acid (OA), were determined. The water content (WC) of the shoots was calculated and the OA components were analyzed. Finally, the osmotic adjustment and ion balance traits in the shoots of K. sieversiana were explored. The results showed that the WC of K. sieversiana remained higher than 6 [g g−1 Dry weight (DW)] even under the highest salt or alkali stress. At salinity levels >240 mM, proline concentrations increased dramatically, with rising salinity. We proposed that this was not a simple response to osmotic stress. The concentrations of Na+ and K+ all increased with increasing salinity, which implies that there was no competitive inhibition for absorption of either in K. sieversiana. Based on our results, the osmotic adjustment feature of salt stress was similar to that of alkali stress in the shoots of K. sieversiana. The shared essential features were that the shoots maintained a state of high WC, OA, Na+, K+ and other inorganic ions, accumulated largely in the vacuoles, and betaine, accumulated in cytoplasm. On the other hand, the ionic balance mechanisms under both stresses were different. Under salt stress, K. sieversiana accumulated OA and inorganic ions to maintain the intracellular ionic equilibrium, with close to equal contributions of OA and inorganic ions to anion. However, under alkali stress, OA was the dominant factor in maintaining ionic equilibrium. The contribution of OA to anion was as high as 84.2%, and the contribution of inorganic anions to anion was only 15.8%. We found that the physiological responses of K. sieversiana to salt and alkali stresses were unique, and that mechanisms existed in it that were different from other naturally alkali-resistant gramineous plants, such as Aneurolepidium chinense, Puccinellia tenuiflora. Responsible Editor: John McPherson Cheeseman.  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