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 187 毫秒
1.
We have studied the movements of H+ from the in vitro frog skin into the outside solution because it has been suggested that the movement of sodium from the outside solution into the skin may result from the forced exchange of Na+ by H+.Our main observations can be summarized as follows: (a) Hydrogen moves from the skin into the outside solution at a rate of 0.04 μequiv · cm?2 · h?1 while Na+ influx had a value of 0.49 μequiv · cm?2 · h?1. (b) The rate of H+ secretion is not significantly affected by substituting the Na+ in the outside solution by K+ nor by inhibiting Na+ influx with amiloride (5 · 10?5 M). (c) Acetazolamide (5 · 10?3 M) blocked H+ secretion without altering the potential difference across the skin. (d) The rate of H+ production is not underestimated because it may have been neutralized by HCO3? secreted into the outside solution in exchange for Cl?. Substituting all the Cl? by SO42? in the outside solutions does not result in an increase in the rate of H+ production. (e) The steady-state rate of H+ secretion is not affected by large changes in electrochemical potential gradients for H+. Neither abolishing the potential difference across the skin nor a 10-fold change in H+ concentration in the outside solution affected significantly the steady-state rate of H+ secretion. (f) The H+ secretion was abolished by the metabolic inhibitors dinitrophenol (1 · 10?4 M) and Antimycin A (1.5 · 10?6 M) which also markedly reduced the potential difference across the skin.Observations (a), (b), and (c) suggest that H+ and Na+ movements across the outer border of the isolated frog skin are not coupled. The ratio of Na+ to H+ movements is very different from unity and Na+ movements can be abolished without any effects on H+ secretion and conversely H+ movements can be abolished without interruption of Na+ uptake.A second conclusion suggested by these results is that the H+ secretion does not result from movement of H+ following its electrochemical potential gradient since that rate of secretion is not affected by marked changes in either potential or [H+]. Furthermore, the effects of metabolic inhibitors suggest that H+ secretion requires the expenditure of energy by the cell.  相似文献   

2.
The preparation of the planar yellow [Ni([8]aneN2)2](ClO4)2 is described. The complex dissociates in basic solution, with rate = kOH[NiL][OH?] (L = 1,5-diazacyclo-octane). At 25 °C, kOH = 4.5 x 10?2 M?1 s?1 and the corresponding activation parameters are ΔH = 69.2 kJ mol?1 and ΔS298 = ?38.6 J K?1 mol?1. Acid catalysed dissociation in quite slow even in strongly acidic solutions. The kinetic data in this case can be fitted to the expression Kobs = ko + KH[H+], where ko relates to a solvolytic pathway and kH to the acid catalysed pathway. At 60 °C, Ko = 2 x 10?5 s?1 and kH is 2 x 10?5 M?1 s?1. Possible mechanisms for these reactions are considered.The Ni(II)/Ni(III) redox couple for NiLn+ is irreversible on Pt using MeCN as solvent.  相似文献   

3.
Abstract Effects of temperature on the ionic relations and energy metabolism of Chara corallina were investigated. Measurements were made of the ionic content, tracer ion fluxes, and photosynthetic and dark CO2 fixation in isolated cells, and of O2 exchange in photosynthesis and respiration in isolated shoot apices. The total intracellular concentration of K+, Na+ and Cl? was the same in cells held for 5 days in non-growing medium at 15°C (the growth temperature) as in those held at 25°C or 5°C. The tracer influx in the light of all ions tested (Rb+, Na+, CH3NH3+, Cl? and H2PO4?) was lower at 5°C than at 15°C in experiments in which cells were subjected to 5°C for less than 24 h in toto. The influx at 25°C was greater than that at 15°C for H2PO?4, there was no difference between the two temperatures for Na+, while the influx at 25°C was less than that at 15°C for Cl?, Rb+ and CH3NH3+ For Cl? and H2PO?4 similar results were found in later experiments with cells grown at 20—23°C. Photosynthetic CO2 fixation and O2 evolution, and respiratory O2 uptake, are greater at 25°C, and lower at 5°C, than they are at the growth temperature of 15°C. In longer-term pretreatments at the different temperatures, tracer Cl? influx at 15°C and particularly at 25°C were lower than in short-term experiments, while the influx at 5°C was higher. It was concluded from these experiments, and from previous data on H+ free energy differences across the plasmalemma, that (1) the maintenance of internal ion concentrations involves a close balancing of influx and efflux of K+, Na+ and Cl? at all experimental temperatures; (2) the regulation of the tracer fluxes of the ions is kinetic rather than thermodynamic and (3) that the tracer fluxes at low temperatures are not restricted by the rate at which respiration or photosynthesis can supply energy to them.  相似文献   

4.
The absorbance change of the weak base dye probe, Acridine orange, was used to monitor alterations of pH gradients across renal brush border membrane vesicles. The presence of Na+/H+ or Li+/H+ exchange was demonstrated by diluting Na2SO4 or Li2SO4 loaded vesicles into Na+- or Li+-free solutions, which caused dye uptake. About 20% of the uptake was abolished by lipid permeable cations such as valinomycin-K+ or tetraphenylphosphonium, indicating perhaps the presence of a finite Na+ conductance smaller than electroneutral Na+/H+ exchange. The protonophore tetrachlorosalicylanilide raised the rate of dye uptake under these conditions, hence the presence of an Na+ conductance greater than the H+ conductance was suggested. K+ gradients also induced changes of pH, at about 10% of the Na+ or Li+ rate. Partial inhibition (21%) was seen with 0.1 mM amiloride indicating that K+ was a low affinity substrate for the Na+/H+ exchange. Acceleration both by tetrachlorosalicylanilide (2-fold) and valinomycin (4-fold) suggested the presence of 2 classes of vesicles, those with high and those with low K+ conductance. The larger magnitude of the valinomycin dependent signal suggested that 75% of the vesicles had a low K+ conductance. Inward Cl? gradients also induced acidification, partially inhibited by the presence of tetraphenylphosphonium, and accelerated by tetrachlorosalicylanilide. Thus both a Cl? conductance greater than the H+ conductance and a Cl?/OH? exchange were present. The rate of Na+/H+ exchange was amiloride sensitive with a pH optimum of 6.5 and an apparent Km for Na+ or Li+ of about 10 mM and an EA of 14.3 kcal per mol. A 61-fold Na2SO4 gradient resulted in a pH gradient of 1.64 units which increased to 1.8 with gramicidin. An equivalent NaCl gradient gave a much lower ΔpH even in the presence of gramicidin showing that the H+ and Cl? pathways could alter the effects of the Na+/H+ exchange.  相似文献   

5.
The effect of the addition of ouabain to the nutrient solution was determined on resistance, potential difference (p.d.) and H+ secretion rate. In NaCl media, 10?3 M ouabain decreased significantly the p.d. from 25.6 mV to 16.1 mV in 30 min and to 11.0 mV in 60 min. NO significant changes occurred in resistance and H+ secretion rate. In Na2SO4 (Cl?-free) media, ouabain produced a biphasic effect on p.d. The p.d. changed from ?28.0 mV (nutrient-negative) to a nadir of ?37.4 mV in 7 min and then increased to ?16.4 mV in 60 min. At the nadir there was no significant change in resistance or H+ secretion rate but at 60 min, unlike Cl? media, resistance increased by 36% and H+ secretion rate decreased by 43%. To decide whether the ouabain-caused decrease in H+ rate in Na2SO4 media was due to an effect on the H+ pump or on resistance of the return pathways, the voltage was clamped at 0 and 40 mV. Clamping the voltage showed that in the case of a marked decrease in the H+ secretion rate, the H+ transport mechanism itself was inhibited (and not the parallel pathway). The decrease in p.d. due to ouabain in Cl? and SO42? media indicates that the (Na+ + K+)-ATPase mechanism may be electrogenic.  相似文献   

6.
H+ extrusion by the isolated skins of two amphibia, Rana ridibunda and Bufo bufo was studied in order to test for the presence of exchange mechanisms of the type Na+/H+ and Cl?/HCO3?, which have been described in several epithelial structures. The preparations were mounted in chambers of the Ussing type, so that the short-circuit current could be used as a function of Na+ transport and the pH-stat technique was utilized to determine the rates of H+ extrusion under different experimental conditions. These conditions were either the withdrawal of the ions intervening in the mentioned exchanges (Cl- or Na+, or the addition of drugs with well-known effects on Na+ uptake and transport (antidiuretic hormone and amiloride).In the frog skin, H+ excretion was detected in solutions containing either Cl? or SO42?, with identical rates. Again, Na+ substitution by Mg2+ had no effect on H+ excretion rates, neither did the suppression of Na+ influx by amiloride or its stimulation by antidiuretic hormone. These experiments were repeated with similar results in gland-free preparations of the epidermis of frog skin separated from the corion by the action of collagenase.Experiments in toad skin showed that H+ excretion could not be detected when Cl? was present in the outer medium, but became apparent if an impermeant anion, SO42?, was used. This observation is compatible with the existence of an exchange mechanism of the type Cl?/HCO3?. Secondly, in these preparations H+ extrusion increased after stimulation with antidiuretic hormone and decreased when amiloride was used or when Na+ was substituted by Mg2+, suggesting that at least a fraction of the total H+ efflux is linked to Na+ influx. In the isolated frog skin this mechanism does not seem to be operative.  相似文献   

7.
Extraction of Ca2+ from the oxygen-evolving complex of photosystem II (PSII) in the absence of a chelator inhibits O2 evolution without significant inhibition of the light-dependent reduction of the exogenous electron acceptor, 2,6-dichlorophenolindophenol (DCPIP) on the reducing side of PSII. The phenomenon is known as “the decoupling effect” (Semin et al. Photosynth Res 98:235–249, 2008). Extraction of Cl? from Ca2+-depleted membranes (PSII[–Ca]) suppresses the reduction of DCPIP. In the current study we investigated the nature of the oxidized substrate and the nature of the product(s) of the substrate oxidation. After elimination of all other possible donors, water was identified as the substrate. Generation of reactive oxygen species HO, H2O2, and O 2 ·? , as possible products of water oxidation in PSII(–Ca) membranes was examined. During the investigation of O 2 ·? production in PSII(–Ca) samples, we found that (i) O 2 ·? is formed on the acceptor side of PSII due to the reduction of O2; (ii) depletion of Cl? does not inhibit water oxidation, but (iii) Cl? depletion does decrease the efficiency of the reduction of exogenous electron acceptors. In the absence of Cl? under aerobic conditions, electron transport is diverted from reducing exogenous acceptors to reducing O2, thereby increasing the rate of O 2 ·? generation. From these observations we conclude that the product of water oxidation is H2O2 and that Cl? anions are not involved in the oxidation of water to H2O2 in decoupled PSII(–Ca) membranes. These results also indicate that Cl? anions are not directly involved in water oxidation by the Mn cluster in the native PSII membranes, but possibly provide access for H2O molecules to the Mn4CaO5 cluster and/or facilitate the release of H+ ions into the lumenal space.  相似文献   

8.
Eu(NO3)3?5H2O and EuCl3?6H2O were allowed to react with bis(diphenylphosphino)alkane dioxides Ph2P(O)(CH2)nP(O)Ph2 (n = 2, 4, 6) to obtain polymeric and binuclear complexes. The prepared compounds were structurally characterized by X‐ray diffraction. Luminescence measurements (emission and excitation spectra, quantum yields, lifetimes) were compared with crystallographic data in order to find a relationship between luminescent properties of the Eu(III) complexes and their structures. The Eu(III) polymers, especially [Eu(dpphO2)2Cl2]+Cl}n, have shown extremely long luminescence lifetimes, up to 3.73 ms, as a result of a highly protecting hydrophobic shield. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

9.
Atmospheric inputs of elements/ions into the soil through bulk precipitation and throughfall (precipitation below tree canopies) were monitored monthly at two forested catchments (Lesni Potok and Liz) in central and southwestern Bohemia, respectively. The annual deposition fluxes (expressed in μg/mg m?2 yr?1) of Al, As, Ba, Be, Ca, Cd, Cl?, F?, Fe, K, Mg, Mn, Ntot, Na, Ni, Pb, Rb, SO 4 2? , Sr and Zn between 1997 and 2005 were calculated from their concentrations in monthly collected samples of both precipitation types. The flux of H+ was calculated from the monthly pH values as well. The more pristine character of the Liz catchment was manifested in lower inputs of anions of strong inorganic acids (mostly of anthropogenic origin) and of H+ in spite of higher precipitation amounts at the site. The comparison of fluxes in bulk precipitation (BP) and throughfall (TH) has shown significantly higher values for Rb, K, Mg, Mn, F?, Ca, SO 4 2? , Sr, Ba and Cl? in the latter flux. It is declared that high fluxes of these elements/ions in TH significantly affect the forest soil water chemistry and that the forest vegetation significantly contributes to the mobilization of several elements in soil and to their redistribution throughout the soil profile.  相似文献   

10.
197Au Mössbauer spectra of the series of complexes of gold(I), Au(nucl)2Cl and gold(III), Au(nucl)Cl3, Au(nucl - H+)Cl2 and Au(nucl)2Cl3 were measured at 4.2 K, (nucl = nucleoside, e.g. guanosine(guo), inosine(ino), triacetylguanosine-(trguo) and triacetylinosine(trino)). It is concluded from the spectra that the gold(I) nucleosides have linear ClAuN coordination, with one coordinated nucleoside molecule per gold(I) ion, bound via the N(7) atom. The σ-donor strength of the guo ligand is somewhat higher than that of the ino ligand. The complexes Au(ino)Cl3 and Au(guo)Cl3, in the series Au(nucl)Cl3, have significantly higher IS and QS values than the corresponding complexes with the triacetylnucleosides, Au(trino)Cl3 and Au(trguo)Cl3. This may be explained by a weak O(6)-interaction with gold(III), in a nearly trigonal bipyramidal configuration in the former case and by the presence of the strongly electron withdrawing acetyl groups in the latter, which reduces the donor strength of their N(7) atoms. The complexes of the Au(nucl - H+)Cl2 series all appear to have a polymeric structure. The gold(III) ion is bound to the N(7) atom and the O(6) or the N(1) atom of the nucleosides. Finally, the Mössbauer spectra of the series Au(nucl)2)Cl3 can only be explained by assuming approximately octahedral AuN2Cl4 structures, with bridging chlorine atoms.  相似文献   

11.
Ascidia callosa sperm are triggered to undergo initiation of the sperm reaction (mitochondrial swelling) by increasing the pH or lowering the Na+ concentration of the medium. The optimal [Na+] for acid release is 20 mM with excellent correlation between acid release and initiation of morphological changes. Increasing the [K+] to around 20 mM inhibits acid release when applied up to 1 min after triggering the sperm but with less inhibition at 2 and 4 min, suggesting that K+ inhibits initiation of acid release rather than acid release itself. Acid release and the sperm reaction can also be triggered by Cl?-free (NO?3 or glutamate substituted) seawater (SW). Cl? efflux accompanies H+ efflux with twice as many Cl? being released as H+. Both H+ and Cl? release in Cl?-free SW are dependent upon CO2 being present in HCO?3-free medium, suggesting that H+ efflux is in part Cl? and HCO?3-mediated. However, the chloride channel blocking agent SITS has no effect on H+ release and augments Cl? release. Acid release results in a substantial increase in internal pH as determined by partitioning of 9-amino acridine. We envision acid release from ascidian sperm as involving two systems, the Na+-dependent acidification system of unreacted sperm and the Cl?- and HCO?3-mediated H+ release at activation. The mechanism controlling acid release would then involve inactivation of the internal acidification process and activation of the chloride-bicarbonate-mediated alkalinization process.  相似文献   

12.
The structures of bis(1H+,5H+-S-methylisothiocarbonohydrazidium) di-μ-chlorooctachlorodibismuthate(III) tetrahydrate: (C2H10N4S)2(Bi2Cl10)· 4H2O (compound [I]) and of tris(1H+-S-methylisothiocarbonohydrazidium) esachlorobismuthate(III): (C2H9N4S)3(BiCl5.67I0.33) (compound [II]) were determined from single crystal X-ray diffractometer data. Both compounds crystallize as triclinic (P ); crystals [I] with Z = 1 formula unit in a cell of constants: A = 10.621(3), B = 9.989(5), C = 7.439(3) Å, α = 88.31(2), β = 84.51(2), γ = 68.88(2)°, final R = 0.0427 for 2229 unique reflections with I 2σ(I); crystals [II] with Z = 2 and cell dimensions: A = 14.109(4), B = 12.209(9), C = 8.206(7) Å, α = 103.54(3), β = 104.95(2), γ = 81.96(2)°, final R = 0.0411 for 3637 unique reflections (1 2σ(I)). The structure of [I] is built up of diprotonated organic cations, water molecules and dinuclear centrosymmetric [Bi2Cl10]4− anions held together by N-HCl, N-HO, O-HCl hydrogen bonds and Van der Waals interactions. The [Bi2Cl10]4− complex consists of two edge-sharing octahedra in which three pairs of bonds of similar length are observed (Bi-Clav = 2.602(5), 2.712(4), 2.855(5) Å). The structure of [II] consists of monoprotonated cations and [BiCl5.67I0.33]3− anions held together by a tridimensional network of hydrogen bonds. Each bismuth atom is octahedrally surrounded by six chlorine atoms, one of which is statistically substituted by a iodine atom.  相似文献   

13.
To evaluate the biological preference of [Yb(phen)2(OH2)Cl3](H2O)2 (phen is 1,10-phenanthroline) for DNA, interaction of Yb(III) complex with DNA in Tris–HCl buffer is studied by various biophysical and spectroscopic techniques which reveal that the complex binds to DNA. The results of fluorescence titration reveal that [Yb(phen)2(OH2)Cl3](H2O)2 has strongly quenched in the presence of DNA. The binding site number n, apparent binding constant K b, and the Stern–Volmer quenching constant K SV are determined. ΔH 0, ΔS 0, and ΔG 0 are obtained based on the quenching constants and thermodynamic theory (ΔH 0?>?0, ΔS 0?>?0, and ΔG 0?<?0). The experimental results show that the Yb(III) complex binds to DNA by non-intercalative mode. Groove binding is the preferred mode of interaction for [Yb(phen)2(OH2)Cl3](H2O)2 to DNA. The DNA cleavage results show that in the absence of any reducing agent, Yb(III) complex can cleave DNA. The antimicrobial screening tests are also recorded and give good results in the presence of Yb(III) complex.  相似文献   

14.
The brown algae Desmarestia ligulata var. ligulata (Lightf.) Lamour., and D. viridis (Mull.) Lamour., accumulate H2SO4 until their average internal pH is 0.5 to 0.8. A related species, D. aculeata (L.) Lamour., does not accumulate acid. The H2SO4 accumulation is accompanied by a reduction in the K+ and Cl content, presumedly to maintain osmotic balance. Measurements of the membrane potential and H+ and SO42− concentrations indicate that both ions are accumulated in the vacuole against their electrochemical potential gradients.

The internal pH remains constant in all three species over the growing season, despite striking changes in the algal morphology. The pH is not affected by periods of darkness of up to 34 hours. Sulfate accumulated in the vacuoles appears to be trapped there since incubation of D. ligulata for up to 10 days in sulfate-free medium resulted in little loss of either vacuolar sulfate or H+. Although the uptake of H2SO4 into the vacuole must require energy, the maintenance of the vacuolar H2SO4 may be due to the impermeability of the tonoplast, with little necessity for continued expenditure of energy.

  相似文献   

15.
Synechococcus R-2 (PCC 7942) actively accumulated Cl? in the light and dark, under control conditions (BG-11 media: pHo, 7·5; [Na+]o, 18 mol m?3; [Cl?]o, 0·508 molm?3). In BG-11 medium [Cl?], was 17·2±0·848 mol m?3 (light), electrochemical potential of Cl? (ΔμCl?i,o) =+211±2mV; [Cl?]i= 1·24±0·11 mol m?3(dark), ΔμCl?i,o=+133±4mV. Cl? fluxes, but not permeabilities, were much higher in the light: ?Cl?i,o= 4·01±5·4 nmol m?2 s?1, PCl?i,o= 47±5pm s?1 (light); ?Cl?i,o= 0·395±0·071 nmol m?2 s?1, PCl?i,o= 69±14 pm s?1 (dark). Chloride fluxes are inhibited by acid pHo (pHo 5; ?Cl?i,o= 0·14±0·04 nmol m?2 s?1); optimal at pHo 7·5 and not strongly inhibited by alkaline pHo (pHo 10; ?Cl?1i,o= 1·7±0·14 nmol m?2 s?1). A Cl?in/2H+in coporter could not account for the accumulation of Cl? alkaline pHo. Permeability of Cl? is very low, below 100pm s?1 under all conditions used, and appears to be maximal at pHo 7·5 (50–70 pm s?1) and minimal in acid pHo (20pm s?1). DCCD (dicyclohexyl-carbodiimide) inhibited ?Cl?i,o in the light about 75% and [Cl?]i fell to 2·2±0·26 (4) mol m?3. Valinomycin had no effect but monensin severely inhibited Cl? uptake ([Cl?]i= 1·02±0·32 mol m?3; ?Cl?i,o= 0·20±0·1 nmol m?2 s?1). Vanadate (200 mmol m?3) accelerated the Cl? flux (?Cl?i,o= 5·28±0·64 nmol m?2 s?1) but slightly decreased accumulation of Cl? ([Cl?], = 13·9±1·3 mol m?3) in BG-11 medium but had no significant effect in Na+-free media. DCMU (dichlorophenyldimethylurea) did not reduce [Cl?], or ?Cl?i,o to that found in the dark ([Cl?]i= 8·41±0·76 mol m?3; ?Cl?i,o= 2·06±0·36 nmol m?2 s?1). Synechococcus also actively accumulated Cl? in Na+-free media, [Cl?]i was lower but ΔΨi,o hyperpolarized in Na+-free media and so the ΔμCl?i,o was little changed ([Cl?]i= 7·98±0·698 mol m?3; ΔμCl?i,o=+203±3 mV). Net Cl? uptake was stimulated by Na+; Li+ acted as a partial analogue for Na+. Synechococcus has a Na+ activated Cl? transporter which is probably a primary 2Cl?/ATP pump. The Cl? pump is voltage sensitive. ΔμCl?i,o is directly proportional to ΔΨi,o(P»0·01%): ΔμCl?i,o= -1·487 (±0·102) ×ΔΨi,o, r= -0·983, n= 31. The ΔμCl?i,o increased (more positive) as the Δμi,o became more negative. The ΔμCl?i,o has no known function, but might provide a driving force for the uptake of micronutrients.  相似文献   

16.
The neutral, noncyclic, imide and ether containing ionophore AS701, has been developed as Li+-selective molecule, to be used potentially as an aid in the Li+-therapy of manic-depressive illness. The present report is a characterization of this molecule in neutral lipid bilayer membranes. This ionophore was found to the bilayers Li+-selective, acting as a selective carrier of monovalent cations. In addition, this molecule was found to be capable of acting as a selective carrier of monovalent anions. For both types of ions, the rate-limitting step in the process of permeation was found to be the diffusion of the carrier-ion complex through the membrane. The membrane-permeating species were found to be 2 : 1 carrier-ion complexes, carrying either a monovalent cation or a monovalent anion. The selectivity sequences among the ions studied being: Li+(1) > ClO4?(0.7) > Na+(0.07) > K+(0.016) > Rb+(0.0095) > Cs+(0.0083) > Cl?(0.001). Mg2+ and SO42? were found to be impermeant (under present experimental conditions). This sequence shows that the AS701 molecule has low selectivity for ions present in biological media, among those studied (i.e. Na+, K+, Mg2+, Cl2? and SO42?). This indicates that these ions will not interfere in the Li+ permeability induced by this carrier in vivo, and that the carrier will not interfere in the normal transport processes of these ions.  相似文献   

17.
The kinetics of the light-driven Cl? uptake pump of Synechococcus R-2 (PCC 7942) were investigated. The kinetics of Cl? uptake were measured in BG-11 medium (pHo, 7·5; [K+]o, 0·35 mol m?3; [Na+]o, 18 mol m?3; [Cl?]o, 0·508 mol m?3) or modified media based on the above. Net36Cl? fluxes (?Cl?o,i) followed Michaelis-Menten kinetics and were stimulated by Na+ [18 mol m?3 Na+ BG-11 ?Cl?max= 3·29±0·60 (49) nmol m?2 s?1 versus Na+-free BG-11 ?Cl?max= 1·02±0·13 (54) nmol m?2 s?1] but the Km was not significantly different in the presence or absence of Na+ at pHo 10; the Km was lower, but not affected by the presence or absence of Na+ [Km = 22·3±3·54 (20) mmol m?3]. Na+ is a non-competitive activator of net ?Cl?o,i. High [K+]o (18 mol m?3) did not stimulate net ?Cl?o,i or change the Km in Na+-free medium. High [K+]o (18 mol m?3) added to Na+ BG-11 medium decreased net ?Cl?o,i [18 mol m?3K+ BG-11; ?Cl?max= 2·50±0·32 (20) nmol m?2 s?1 versus BG-11 medium; ?Cl?max= 3·35±0·56 (20) nmol m?2 s?1] but did not affect the Km 55·8±8·100 (40) mmol m?3]. Na+-stimulation of net ?Cl?o,i followed Michaelis-Menten kinetics up to 2–5 mol m?3 [Na+]o but higher concentrations were inhibitory. The Km for Na+-stimulation of net ?Cl?o,i [K1/2(Na+)] was different at 47 mmol m?3 [Cl?]o (K1/2[Na+] = 123±27 (37) mmol m?3]. Li+ was only about one-third as effective as Na+ in stimulating Cl? uptake but the activation constant was similar [K1/2(Li+) = 88±46 (16) mmol m?3]. Br? was a competitive inhibitor of Cl? uptake. The inhibition constant (Ki) was not significantly different in the presence and absence of Na+. The overall Ki was 297±23 (45) mmol m?3. The discrimination ratio of Cl? over Br? (δCl?/δBr?) was 6·38±0·92 (df = 147). Synechococcus has a single Na+-stimulated Cl? pump because the Km of the Cl? transporter and its discrimination between Cl? and Br? are not significantly different in the presence and absence of Na+. The Cl? pump is probably driven by ATP.  相似文献   

18.
A variety of metal ions can bind to the iron-transport protein, transferrin, at two specific sites. For each metal ion, a carboxylate anion is concomitantly bound. Six metal ions which were examined fall into two classes based on proton release and ultraviolet spectral changes which accompany binding to the protein. Class II ions, which include Cu2+ and Zn2+, release approximately 2 H+/metal bond. Class III ions, which include Fe3+, Ga3+, Al3+, and VO2+, release approximately 3 H+/metal bound. The increase in absorbance near 242 nm, characteristic of tyrosine ionization, has the ratio 0.55–0.75 for class II:class III ions. Both Fe3+ and Cu2+ form metal-transferrin-oxalate complexes in the presence of excess C2O42?. Fe3+ releases close to 3 H+/metal whether forming oxalate or bicarbonate complexes with transferrin. Binding of Cu2+ to transferrin releases 2 H+/metal in the presence of C2O2?4 or HCO3?. Since equal numbers of H+/metal are released for both anions, it is likely that the bicarbonate ion does not lose its proton, and remains as HCO3? in transferrin. These results are interpreted in terms of possible combinations of ligands at the metal binding sites.  相似文献   

19.
Biomineralization on bacterial surface is affected by biomolecules of bacterial cell surface. Lipopolysaccharide (LPS) is the main and outermost component on the extracellular membrane of Gram-negative bacteria. In the present study, the molecular mechanism of LPS in affecting biomineralization of Ag+/Cl? colloids was investigated by taking advantages of two LPS structural deficient mutants of Escherichia coli. The two mutants were generated by impairing the expression of waaP or wbbH genes with CRISPR/Cas9 technology and it induced deficient polysaccharide chain of O-antigen (ΔwbbH) or phosphate groups of core oligosaccharide (ΔwaaP) in LPS structures. There were significant changes of the cell morphology and surface charge of the two mutants in comparing with that of wild type cells. LPS from ΔwaaP mutant showed increased ΔHITC upon interacting with free Ag+ ions than LPS from wild type cells or ΔwbbH mutant, implying the binding affinity of LPS to Ag+ ions is affected by the phosphate groups in core oligosaccharide. LPS from ΔwbbH mutant showed decreased endotherm (ΔQ) upon interacting with Ag+/Cl? colloids than LPS from wild type or ΔwaaP mutant cells, implying LPS polysaccharide chain structure is critical for stabilizing Ag+/Cl? colloids. Biomineralization of Ag+/Cl? colloids on ΔwbbH mutant cell surface showed distinctive morphology in comparison with that of wild type or ΔwaaP mutant cells, which confirmed the critical role of O-antigen of LPS in biomineralization. The present work provided molecular evidence of the relationship between LPS structure, ions, and ionic colloids in biomineralization on bacterial cell surface.  相似文献   

20.
Rate parameters have been obtained for the oxidation of cuprous stellacyanin by cobalt(III) ions of the form cis(N)-[CoN2O4]?, including cis(N)-[Co(NTA)(gly)]?, cis(N)-[Co(IDA)2]?, [Co(en)(ox)2]?(μ 0.5 M(phosphate), pH 7.0), and Co(EDTA)?(μ 0.1 M(NaCl), pH 7.2, 0.001 M phosphate). An excellent isokinetic correlation between the activation parameters ΔH and ΔS exists for the reactions of aminopolycarboxylatocobalt(III) ions with reduced stellacyanin (β = 300 ± 12 K; correlation coefficient = 0.995). It is concluded that enthalpy-entropy compensation in these reactions may be understood in terms of differing orientations preferred by the various oxidants in forming precursor complexes with the reduced blue protein. While ΔH and ΔS values for electron transfer from stellacyanin to cis(N)-[CoN2O4]? ions vary over ranges of 10.7 kcal/mol and 34 cal/mol-deg, respectively, room temperature rate constants are relatively constant (3.6–34.5 M?1 sec?1), as expected from Marcus theory for outer sphere electron transfer.  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