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 pH dependence of the initial uptake of norepinephrine by rat whole brain synaptosomes was studied using short incubation times at 37 degrees C in order to examine the possible involvement of the phenolic OH group. The pH vs. uptake profile exhibits a maximum near pH 8.2 in H2O medium. When the medium was changed to 2H2O, the profile showed a shift of maximum corresponding to the pKa change of the phenolic OH group. The pH vs. uptake profile of tyramine was quite different from that of norepinephrine. These pH effects on uptake were explained as manifestations of the involvement of the phenolic OH group in the process. The amine and phenolic hydroxyl groups in norepinephrine were studied separately by employing two series of compounds structurally related to catecholamines, amphetamine-like and catechol-like, for their inhibitory effects on the uptake. The inhibitions were affected by changes in pH with changes in opposite directions found for the two series indicating the need for a positive charge in the side chain and suggesting an effect of the negative charge on the ring. These charge characteristics agreed with the pH profile observed in uptake. Consequently, the two groups with opposite charge characteristics in norepinephrine both appear to function in the uptake process.  相似文献   

2.
The thermal perturbation difference spectra of phenolic and indolic chromophores in water resemble the isothermal D2O and H2O spectra of these chromophores. For phenols approximately equal Δ? values are obtained in both types of spectra, but for their methyl ethers Δ? values of D2O vs H2O spectra are about half of those of the thermal perturbation spectra. Phenols and their methyl ethers were studied in deuterated ethylene glycol and glycerol vs the corresponding protiated solvent, and in nonprotic solvents containing 0.25–4% D2O or H2O. For phenols in D2O vs H2O, about one-third to one-half of the difference spectrum is attributed to solvent structure difference, and the remainder to the effects of replacing OH by OD and to differences in accepting hydrogen bonds from D2O and H2O. The refractive index difference between D2O and H2O was shown to be a minor contribution by means of experiments in which D2O was at 5 dgC and H2O at 47 dgC, conditions of equal refractive index (NaD). D2O vs H2O and glycerol-d vs glycerol-h difference spectra of ribonuclease are about twice as large as expected from the known number of exposed tyrosyl side chains. Possible sources of error in D2O vs H2O spectra of proteins are discussed.  相似文献   

3.
The crystal structure of a complex of α-cyclodextrin (α-CD) with 2-fluoro-4-nitrophenol · 3H2O has been determined by the X-ray diffraction technique. The complex crystallizes in space group P212121 with cell dimensions: a = 13.431(3), b = 15.299(4), c = 24.780(5) Å. The structure was solved by direct methods and refined to R = 6.7% for 4483 reflections. The crystal structure is isomorphous to the α-CD-4-nitrophneol · 3H2O complex. The phenyl group is inside the cavity, so that the O-4 hexagon of the α-CD is distorted in a systematic manner: the longest diagonal [O-4(G2) O-4(G5)] is in the direction of the benzene ring. The phenolic OH group protrudes from the secondary OH side of the cavity and the NO2 group is situated on the primary OH side. The hydrophobic F atom is statitically disordered over two sites and is located in the hydrophilic space, just beyond the rim of the secondary OH side of the cavity.  相似文献   

4.
《BBA》1986,848(1):1-7
An almost stoichiometric conversion of nitrite to nitrous oxide was observed during the nitrite reduction by Paracoccus denitrificans cells in a medium of pH 6.4. The N2O accumulated in the reaction medium and was decomposed only after nitrite had been consumed; when the pH of the medium was higher than 7.3–7.4, nitrous oxide did not accumulate. The activity of N2O reductase was, in the whole range of pH 6.4–9.2, higher than the activity of NO2 reductase, both activities showing the maximum at the pH higher than 8.0. Using an artificial donor, TMPD plus ascorbate, the maximum activity of NO2 reductase, but not N2O reductase was shifted by about two pH units to acidic region. The activity of nitrite reductase declined in the presence of N2O only at higher pH values. Cytochrome c, as a common electron donor for both N2O and NO2 reductase, was more oxidized at pH < 7.3 in the presence of NO2 than in the presence of N2O, the opposite being true at pH > 7.3. The increased flux of electrons to cytochrome c has for a constant pH value (6.4) no effect on their distribution over NO2 and N2O. The results indicate that the distribution of electrons in the terminal part is determined by the different pH optima for NO2 reductase and N2O reductase, and by a mutual dependence of activities of the two reductases due to the competition for redox equivalents from a substrate.  相似文献   

5.
Ethanol and water are the solvents most commonly used to extract flavonoids from propolis. Do hydrogen-bonding interactions exist between flavonoids and ethanol/water? In this work, this question was addressed by using density functional theory (DFT) to provide information on the hydrogen-bonding interactions between flavonoids and ethanol/water. Chrysin and Galangin were chosen as the representative flavonoids. The investigated complexes included chrysin–H2O, chrysin–CH3CH2OH, galangin–H2O and galangin–CH3CH2OH dyads. Molecular geometries, hydrogen-bond binding energies, charges of monomers and dyads, and topological analysis were studied at the B3LYP/M062X level of theory with the 6?31++G(d,p) basis set. The main conclusions were: (1) nine and ten optimized hydrogen-bond geometries were obtained for chrysin–H2O/CH3CH2OH and galangin–H2O/CH3CH2OH complexes, respectively. (2) The hydrogen atoms except aromatic H1 and H5 and all of the oxygen atoms can form hydrogen-bonds with H2O and CH3CH2OH. Ethanol and water form strong hydrogen-bonds with the hydroxyl, carbonyl and ether groups in chrysin/galangin and form weak hydrogen-bonds with aromatic hydrogen atoms. Except in structures labeled A and B, chrysin and galangin interact more strongly with H2O than CH3CH2OH. (3) When chrysin and galangin form hydrogen-bonds with H2O and CH3CH2OH, charge transfers from the hydrogen-bond acceptor (H2O and CH3CH2OH in structures A, B, G, H, I, J) to the hydrogen-bond donor (chrysin and galangin in structure A, B, G, H, I, J). The stronger hydrogen-bond makes the hydrogen-bond donor lose more charge (A> B> G> H> I> J). (4) Most of the hydrogen-bonds in chrysin/galangin?H2O/CH3CH2OH complexes may be considered as electrostatic dominant, while C?O2···H in structures labeled E and C?O5···H in structures labeled J are hydrogen-bonds combined of electrostatic and covalent characters. H9, H7, and O4 are the preferred hydrogen-bonding sites.  相似文献   

6.
An obligatory anaerobic bacterium was isolated from a mediator-less microbial fuel cell using starch processing wastewater as the fuel and designated as EG3. The isolate was Gram-positive, motile and rod (2.8–3.0 μm long, 0.5–0.6 μm wide). The partial 16S rRNA gene sequence and analysis of the cellular fatty acids profile suggested that EG3 clusters with Clostridium sub-phylum and exhibited the highest similarity (98%) with Clostridium butyricum. The temperature and pH optimum for growth were 37°C and 7.0, respectively. The major products of glucose and glucose/Fe(O)OH metabolism were lactate, formate, butyrate, acetate, CO2and H2. Growth was faster at the initial phase and the cell yield was higher when the medium was supplemented with Fe(O)OH than without Fe(O)OH. These results suggest that Fe(III) ion is utilised as an electron sink. Cyclic voltammetry showed that Clostridium butyricum EG3 cells were electrochemically active. It is a novel characteristic of strict anaerobic Gram-positive bacteria.  相似文献   

7.
C.A. Wraight 《BBA》1979,548(2):309-327
The photoreduction of ubiquinone in the electron acceptor complex (Q1Q11) of photosynthetic reaction centers from Rhodopseudomonas sphaeroides, R26, was studied in a series of short, saturating flashes. The specific involvement of H+ in the reduction was revealed by the pH dependence of the electron transfer events and by net H+ binding during the formation of ubiquinol, which requires two turnovers of the photochemical act. On the first flash Q11 receives an electron via Q1 to form a stable ubisemiquinone anion (Q??11); the second flash generates Q??1. At low pH the two semiquinones rapidly disproportionate with the uptake of 2 H+, to produce Q11H2. This yields out-of-phase binary oscillations for the formation of anionic semiquinone and for H+ uptake. Above pH 6 there is a progressive increase in H+ binding on the first flash and an equivalent decrease in binding on the second flash until, at about pH 9.5, the extent of H+ binding is the same on all flashes. The semiquinone oscillations, however, are undiminished up to pH 9. It is suggested that a non-chromophoric, acid-base group undergoes a pK shift in response to the appearance of the anionic semiquinone and that this group is the site of protonation on the first flash. The acid-base group, which may be in the reaction center protein, appears to be subsequently involved in the protonation events leading to fully reduced ubiquinol. The other proton in the two electron reduction of ubiquinone is always taken up on the second flash and is bound directly to Q??11. At pH values above 8.0, it is rate limiting for the disproportionation and the kinetics, which are diffusion controlled, are properly responsive to the prevailing pH. Below pH 8, however, a further step in the reaction mechanism was shown to be rate limiting for both H+ binding electron transfer following the second flash.  相似文献   

8.
Satter RL  Xu Y  Depass A 《Plant physiology》1987,85(3):850-855
Previous studies reveal that dark-induced closure of Samanea leaflets is accompanied by H+ secretion from flexor motor cells. We now report that flexor tissue excised in the light, incubated in a weakly buffered bathing solution, and then darkened at different temperatures (18°C-30°C) acidified the medium (indicating net H+ efflux) at all temperatures tested, but most rapidly at the highest temperature. However, pH changes reversed direction after 20 to 70 minutes; the lower the temperature, the later pH reversal occurred, and the lower the pH at reversal and after 45 minutes. These data provide a basis for the previously reported promotive effect of low temperature on dark-induced leaflet closure, assuming net H+ and K+ fluxes are opposite in direction. Net H+ efflux at all temperatures tested was greater when the impermeant molecule iminodiacetate replaced small permeant anions in the bathing solution, suggesting that H+ uptake is coupled to anion uptake, probably via a H+/anion symport system. When permeant anions were deficient, the amount of malate in the tissue increased, presumably by new synthesis. Malate synthesis would substitute for H+/anion uptake in charge balance and in providing H+ for cytoplasmic pH regulation.  相似文献   

9.
Spiller H 《Plant physiology》1980,66(3):446-450
Spheroplasts from Anabaena 7119 (formerly designated Nostoc muscorum) were prepared in the presence of serum albumin in 0.5 molar sucrose. Electron transport and photophosphorylation were preserved (> 70% of the maximum rate for 1 week). The pH profile of electron transport and photophosphorylation in the reactions H2O → NADP, H2O → methyl viologen, and H2O → ferricyanide shows that uncoupling by ammonia is small throughout and increases slightly with higher pH. ADP + Pi increased NADP reduction from H2O by 2.5-fold. The ratios of ATP formed per electron pair transported ranged from 0.9 to 1.5. Effects of catalase and superoxide dismutase on the overall O2 balance implicate pseudocyclic electron transport and phosphorylation. The quenching of 9-aminoacridine fluorescence indicates the formation of a Δ pH from 2 to 2.6 during illumination. This pH gradient is abolished by uncouplers; however, complete uncoupling is achieved only by 3-chlorocarbonyl cyanide phenylhydrazone or valinomycin + NH4+. In the presence of NH4+ alone, the membrane potential may act as the driving force for photophosphorylation.  相似文献   

10.
Sakano K 《Plant physiology》1990,93(2):479-483
Upon absorption of phosphate, cultured cells of Catharanthus roseus (L.) G. Don caused a rapid alkalinization of the medium in which they were suspended. The alkalinization continued until the added phosphate was completely exhausted from the medium, at which time the pH of the medium started to drop sharply toward the original pH value. Phosphate exposure caused the pH of the medium to increase from pH 3.5 to values as high as 5.8, while the rate of phosphate uptake was constant throughout (10-17 micromoles per hour per gram fresh weight). This indicates that no apparent pH optimum exists for the phosphate uptake by the cultured cells. The amount of protons cotransported with phosphate was calculated from the observed pH change up to the maximum alkalinization and the titration curve of the cell suspension. Proton/phosphate transport stoichiometry ranged from less than unity to 4 according to the amount of phosphate applied. At low phosphate doses, the stoichiometries were close to 4, while at high phosphate doses, smaller stoichiometries were observed. This suggests that, at high phosphate doses, activation of the proton pump is induced by the longer lasting proton influx acidifying the cytoplasm. The increased H+ efflux due to the proton pump could partially compensate protons taken up via the proton-phosphate cotransport system. Thus, the H+/H2PO4 stoichiometry of the cotransport is most likely to be 4.  相似文献   

11.
(1) The kinetics of isotope exchange catalysed by the membrane-bound hydrogenase of Paracoccus denitrificans have been studied by measuring H2H, H2 or 2H2 produced when the enzyme catalyses the exchange between 2H2 and H2O or H2 and 2H2O. (2) In the 2H2-H2O system the measured rate of H2 production was always higher than that of H2H. The H2H2H ratio remained constant (about 1.70) in the protein concentration range 0.08–1.32 mg. The very rapid formation of H2 with respect to H2H is consistent with the hypothesis of a heterolytic cleavage of 2H2 into a deuteron and an enzyme hydride that can exchange with the solvent. (3) In the H2-2H2O system, the exchange rate was much lower than in the 2H2-H2O system, indicating a marked isotopic effect of 2H2O. (4) The H-2H exchange activity, determined from the initial velocity of H2H formation, is optimal at pH 4.5. A second maximum of activity is observed at pH 8.3. The pH value of 4.5 is also the pH optimum for H2 production while at pH 8.3–8.5 there is a maximum of H2 oxidation activity. (5) In ordinary H2O the Km for hydrogen uptake estimated either from H2 consumption or from benzyl viologen reduction was 0.06–0.07 μM for both H2 and 2H2 indicating a strong affinity of the enzyme for hydrogen at pH 8.3–8.5. Shifting from H2O to 2H2O does not affect the Km of the enzyme for H2 but lowers the Vmax value about 10-fold. The Km for benzyl viologen and methyl viologen was 0.08 and 2 mM, respectively.  相似文献   

12.
The influence of a Donnan effect on the transport of glycine by hemolysed and restored pigeon red cells was examined. The Donnan effect was produced by replacing Cl? with 2,4-toluenedisulfonate or glutamate. The effects of the associated membrane potential and inside-outside pH difference on glycine entry and exit rates were examined. The effects of pH on entry and exit rates in the absence of a Donnan effect were also examined.In the absence of a Donnan effect, Na+-dependent glycine entry requires the protonated form of a group with a pKapp of 7.9 and the depronated form of another group with a pKapp of 6.8. Neither of these are required for exit but the deprotonated form of a group(s) with a pKapp of 6.2 is required. The pK 7.9 group and pK 6.2 group probably react with H+ at the inner face of the membrane and the pK 6.8 group probably reacts at the outer face.The V for glycine entry was determined for cells with their Cl? largely replaced by toluenedisulfonate and without such replacement. Between pH 6.1 and 7, the ratio of the respective V values, VT/VCl, was 1.5–1.7. VT/VCl rose above pH 7 to near 4 at pH 8.3. At pH 6.9, with glutamate replacing cell Cl?, the analogous ratio (VGlu/VCl) was 1.7. The increase of VT/VCl above pH 7 could be quantitatively accounted for by the increase in cell [H+]/medium [H+] caused by the Donnan effect together with the assumption that the pK 7.9 group reacts with H+ at the inner face of the membrane.When cell Cl? was replaced by toluenedisulfonate or glutamate there was a drop in the term in the glycine Km describing Na+ dependence of glycine entry. When cell Cl? was replaced by toluenedisulfonate there was a rise in the Na+-independent term in the glycine entry Km. By replacing varying amounts of cell Cl? with either toluenedisulfonate or glutamate, plots were obtained of entry rates vs. the cell [Cl?]/medium [Cl?] ratio consistent with the assumption that the Donnan-induced membrane potential acts on a “moving” charge. Glycine exit was only slightly accelerated by trans-toluenedisulfonate. The ratio, exit rate into toluenedisulfonate medium/exit rate into Cl? medium rose with decreasing pH. This rise could be accounted for by a Donnan-induced inside-outside pH difference which affects a pKapp 6.2 group reacting with internal H+.The observed influences of the Donnan effect on V(glycine entry), on both components of Km(glycine entry), on the shape of the plot of glycine entry rate vs. the cell [Cl?]/medium [Cl?] ratio and on glycine exit all fit the assumptions that when the empty porter reorients, one unit of negative charge accompanies it “across” the membrane and that no other steps involve charge movement.The properties of the system seem inconsistent with a translational (“ferry boar”) mobile carrier.  相似文献   

13.
Energetically-coupled processes (electron flow, proton uptake and correlated pH gradient) were investigated on envelope-free chloroplasts of lettuce suspended in 1H2O or 2H2O media. Study of the light-intensity and temperature dependencies of these phenomena led to the following observations: 1. At neutral pH, 2H2O diminishes the transmembrane H+ gradient in strong light (chain Photosystem II + Photosystem I) but not in low light; the total H+ uptake is increased at all light intensities: the buffering capacity of the inner compartment is increased in heavy water, possibly through enhancement of interactions between membranous titrable groups and the aqueous phase. 2. 2H2O does not affect the photochemical events of the redox chain, whatever the electron pathway (PSII, PSI or PSII + PSI): only thermal steps are inhibited. The diminution of the apparent quantum yield, sometimes observed, may be ascribed to the dual site of action of the artificial redox carrier (ferricyanide) then used. 3. 2H2O does not modify the activation energy of the limiting step of the electron flow (PSII + PSI) in uncoupled (44 vs. 47 kJ · mol?1) or — but less clearly — in coupled, i.e., ‘basal’, state (55 vs. 59 kJ · mol?1). 2H2O does not either change the temperature of the phase transition of the membrane (17°C) for the uncoupled flow. However, a low-temperature transition, observed only for the coupled chain, is slightly increased by 2H2O; this thermal transition is attributed to the freezing of some bound water near the plastoquinone pool. 4. Δp2H is smaller than Δp1H at all temperatures (PSII + PSI chain). ΔpH is quasi-constant from 0°C to 10°C, then decreases when temperature rises. 2H2O does not change the activation energy of the dark passive H+ efflux, which is almost twice that of the coupled electron flow. The phase transition at low temperature suggests that the proton efflux occurs via two parallel pathways, one temperature-dependent and the other temperature-independent. Except for the increase of the internal buffering capacity, the effects of 2H2O on the membrane conformation seem limited, as shown by the unchanged activation energies of the electron flow and of the H+ leakage. The null activation energy observed at low temperature emphasizes the role of the bound water in these processes; however, the different effects of 2H2O on the transition temperatures indicate that this bound water has different properties when associated with the translocation sites or with the H+ leakage ones. This ‘microcompartmentation’ of the membranes is consistent with the concept of lateral pH heterogeneity we have previously suggested (de Kouchkovsky, Y., and Haraux, F. (1981) Biochem. Biophys. Res. Commun. 99, 205–212). The theoretical computations and the experimental results suggest that in the steady state, the internal pH would be several tenths of a ‘unit’ lower near the plastoquinones than near the H+ efflux sites (coupling factors); this difference would be increased when 2H+ replaces 1H+, owing to the lower mobility of the deuteron. It is concluded that local, and not average, pH (and ΔpH) should be considered for the understanding of the energy transduction processes.  相似文献   

14.
Carbon uptake in the green macroalga Cladophora glomerata (L.) Kütz. from the brackish Baltic Sea was studied by recording changes in pH, alkalinity, and inorganic carbon concentration of the seawater medium during photosynthesis. The use of specific inhibitors identified three uptake mechanisms: 1) dehydration of HCO3 ? into CO2 by periplasmic carbonic anhydrase, followed by diffusion of CO2 into the cell; 2) direct uptake of HCO3 ? via a 4,4′‐diisothiocyanato‐stilbene‐2,2′‐disulfonate‐sensitive mechanism; and 3) uptake of inorganic carbon by the involvement of a vanadate‐sensitive P‐type H + ‐ATPase (proton pump). A decrease in the alkalinity of the seawater medium during carbon uptake, except when treated with vanadate, indicated a net uptake of the ionic species contributing to alkalinity (i.e. HCO3 ? , CO32 ? , and OH ? ) from the medium, where OH ? influx is equivalent to H + efflux. This would suggest that the proton pump is involved in HCO3 ? transport. We also show that the proton pump can be induced by carbon limitation. The inducibility of carbon uptake in C. glomerata may partly explain why this species is so successful in the upper littoral zone of the Baltic Sea. Usually, carbon limitation is not a problem in the upper littoral of the sea. However, it may occur frequently within dense Cladophora belts with high photosynthetic rates that create high pH and low carbon concentrations in the alga's microenvironment.  相似文献   

15.
16.
Reactions of salicyl- and 3,5-dichlorosalicylaldehyde-S-propyl-thiosemicarbazones with salicyl- and 3,5-dichlorosalicylaldehyde in the presence of UO2(CH3COO)2 in different alcohols yielded stable solid complexes corresponding to the general formula [UO2(L)ROH] (R: propyl-, butyl-, pentyl-, and octyl-). The complexes were characterized by means of elemental analysis, IR and 1H NMR spectroscopies. The thermal stabilities of the alcohol solvated complexes were investigated in air and nitrogen atm., and determined their decomposition phases. In the crystal structure of the [UO2(L)(C4H9OH)], the U(VI) centre is seven-coordinated in a distorted pentagonal bipyramidal geometry involving O,O,N,N atoms of two phenolic and two imine groups and one oxygen atom of alcohol molecule in basal plane and two O atoms of dioxo group in apical positions. The title structure is stabilized by one intramolecular interaction of types C-H?Cl and by two intermolecular interactions of types O-H?O and C-H?π (benzene) leading to the molecular chain along the [0 1 0] direction.  相似文献   

17.
The contamination of drinking and irrigation water by arsenic is a severe health risk to millions of people, particularly in developing countries. Arsenic treatment methods therefore need to advance to more durable and cost-effective solutions. In recent years, the unique properties of nanomaterials have received much attention in water treatment research, and their properties (e.g., high number of reactive surface binding sites) may make them suitable for arsenic removal. The aluminum nanoclusters Al13 (AlO4Al12(OH)24H2O12 7+) and Al30 (Al2O8Al28(OH)56(H2O)26 18+) have high specific surface charge, deprotonate over a wide pH range and exhibit a high reactivity due to a great number of OH and H2O groups. This contribution evaluates these chemical properties of aluminum nanoclusters and their efficiency for water treatment, particularly for arsenic removal. It assesses the advantages and constraints when applied in an industrially produced aluminum coagulant or in Al granulate during water treatment.  相似文献   

18.
The composition of taxifolin-iron ions complexes in an octanol-water biphasic system was studied using the method of absorption spectrophotometry. It was found that at pH 5.0 in an aqueous biphasic system the complex of [Tf · Fe2(OH) k (H2O)8 ? k ] is present, but at pH 7.0 and 9.0 the complexes of [Tf2 · Fe(OH) k (H2O)2 ? k ] and [Tf · Fe(OH) k (H2O)4 ? k ] are predominantly observed. The formation of a stable [Tf3 · Fe] complex occurred in octanol phase. The charged iron ion of this complex is surrounded by taxifolin molecules, which shield the iron ion from lipophilic solvent. During transition from water to octanol phase the changes of the composition of complexes are accompanied by reciprocal changes in portion of taxifolin and iron ions in these phases. It was shown that the portion of taxifolin in aqueous solution in the presence of iron ions is increased at high pH values, and the portion of iron ions is minimal at pH 7.0. In addition, the parameters of solubility limits of taxifoliniron ions complexes in an aqueous solution were determined. The data obtained gain a better understanding of the role of complexation of polyphenol with metal of variable valency in passive transport of flavonoids and metal ions across lipid membranes.  相似文献   

19.
The reaction of Cp2TiCl2 with two equivalents of maltol (3-hydroxy-2-methyl-4-pyrone) in water, at room temperature and pH of 5.4, leads to a complete replacement of Cp and chloride ligands affording, Ti(maltolato)2(OH)2. The complex has been characterized by IR, NMR and ESI-MS spectroscopic and cyclic voltammetry methods. In DMSO-d6 solution, the complex shows two isomers in a ratio of 4:1, in which one OH signal can be identified per isomer. This suggests that in solution the complex is monomeric, most likely a chiral cis-Ti(maltolato)2(OH)2 and trans-Ti(maltolato)2(OH)2. The monomeric nature of the complex (in water/methanol 1:1) was verified by ESI-MS spectroscopy, showing a parent peak at 329 m/z. Electrochemical behavior of Ti(maltolato)2(OH)2using cyclic voltammetry experiments showed the complex undergoes irreversible reduction in aprotic solvents. In D2O solution, at pH of 8.4, the 1H NMR spectrum of the complex shows a mixture of monomer and tetramer Ti(IV)-maltol complexes in a ratio of 1:1. The crystallization of Ti(maltolato)2(OH)2 at pH of 8.4 leads to the formation of [Ti4(maltolato)8(μ-O4)] · 18H2O. A single crystal of [Ti4(maltolato)8(μ-O4)] · 18H2O was analyzed by X-ray diffraction methods. Solid state structure determination of the Ti-maltol complex showed to be tetrameric, containing two bridging oxides (in cis position) and two bidentate maltol ligands per titanium in a pseudo-octahedral coordination geometry.  相似文献   

20.
This study was undertaken in order to demonstrate the extent to which the activity of the plasmalemma H+-ATPase compensates for the charge and acidity flow caused by the sugar-proton symport in cells of chlorella vulgaris Beij.. Detailed analysis of H+ and K+ fluxes from and into the medium together with measurements of respiration, cytoplasmic pH, and cellular ATP-levels indicate three consecutive phases after the onset of H+ symport. Phase 1 occurred immediately after addition of sugar, with an uptake of H+ by the hexoseproton symport and charge compensation by K+ loss from the cells and, to a smaller degree, by loss of another ion, probably a divalent cation. This phase coincided with strong membrane depolarization. Phase 2 started approximately 5 s after addition of sugar, when the acceleration of the H+-ATPase caused a slow-down of the K+ efflux, a decrease in the cellular ATP level and an increase in respiration. The increased respiration was most probably responsible for a pronounced net acidification of the medium. This phase was inhibited in deuterium oxide. In phase 3, finally, a slow rate of net H+ uptake and K+ loss was established for several further minutes, together with a slight depolarization of the membrane. There was hardly any pH change in the cytoplasm, because the cytoplasmic buffering capacity was high enough to stabilize the pH for several minutes despite the net H+ fluxes. The quantitative participation of the several phases of H+ and K+ flow depended on the pH of the medium, the ambient Ca2+ concentration, and the metabolic fate of the transported sugar. The results indicate that the activity of the H+-ATPase never fully compensated for H+ uptake by the sugar-symport system, because at least 10% of symport-caused charge inflow was compensated for by K+ efflux. The restoration of pH in the cytoplasm and in the medium was probably achieved by metabolic reactions connected to increased glycolysis and respiration.Abbreviations DMO dimethyloxazolidinedione - EDTA ethylcnediaminetetraacetic acid - p.c. packed cell volume  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