首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
A chloride channel from impermeant sarcoplasmic reticulum (SR) was embedded in a planar lipid bilayer (BLM) and its electrical properties determined. Studies using ER-derived vesicles fused with BLMs have shown that they are permeable to Na+, K+, choline and Cl but less permeable to Ca2+ and Mg2+. Though highly permeable to K+, the liver ER membrane has been postulated to a lack of an efficient ion-conducting structure for K+ channel in the SR. The present study was undertaken with the aim to look at the anionic, Ca2+ and K+ permeability pathways present in the ER membrane. Our reconstituted system exhibits considerable anionic permeability following the sequence: SCN>I>BrCl>gluconate. The findings suggest the chloride channels have low field-strength sites. It can be pharmacologically dissected to Zn2+-sensitive and DIDS-sensitive types. The gating of the channel is weakly voltage-dependent and at higher positive or negative voltages the channel prefers the low sub-conductance states.  相似文献   

2.
Ligand substitution kinetics for the reaction [PtIVMe3(X)(NN)]+NaY=[PtIVMe3(Y)(NN)]+NaX, where NN=bipy or phen, X=MeO, CH3COO, or HCOO, and Y=SCN or N3, has been studied in methanol at various temperatures. The kinetic parameters for the reaction are as follows. The reaction of [PtMe3(OMe)(phen)] with NaSCN: k1=36.1±10.0 s−1; ΔH1=65.9±14.2 kJ mol−1; ΔS1=6±47 J mol−1 K−1; k−2=0.0355±0.0034 s−1; ΔH−2=63.8±1.1 kJ mol−1; ΔS−2=−58.8±3.6 J mol−1 K−1; and k−1/k2=148±19. The reaction of [PtMe3(OAc)(bipy)] with NaN3: k1=26.2±0.1 s−1; ΔH1=60.5±6.6 kJ mol−1; ΔS1=−14±22 J mol−1K−1; k−2=0.134±0.081 s−1; ΔH−2=74.1±24.3 kJ mol−1; ΔS−2=−10±82 J mol−1K−1; and k−1/k2=0.479±0.012. The reaction of [PtMe3(OAc)(bipy)] with NaSCN: k1=26.4±0.3 s−1; ΔH1=59.6±6.7 kJ mol−1; ΔS1=−17±23 J mol−1K−1; k−2=0.174±0.200 s−1; ΔH−2=62.7±10.3 kJ mol−1; ΔS−2=−48±35 J mol−1K−1; and k−1/k2=1.01±0.08. The reaction of [PtMe3(OOCH)(bipy)] with NaN3: k1=36.8±0.3 s−1; ΔH1=66.4±4.7 kJ mol−1; ΔS1=7±16 J mol−1K−1; k−2=0.164±0.076 s−1; ΔH−2=47.0±18.1 kJ mol−1; ΔS−2=−101±61 J mol−1 K−1; and k−1/k2=5.90±0.18. The reaction of [PtMe3(OOCH)(bipy)] with NaSCN: k1 =33.5±0.2 s−1; ΔH1=58.0±0.4 kJ mol−1; ΔS1=−20.5±1.6 J mol−1 K−1; k−2=0.222±0.083 s−1; ΔH−2=54.9±6.3 kJ mol−1; ΔS−2=−73.0±21.3 J mol−1 K−1; and k−1/k2=12.0±0.3. Conditional pseudo-first-order rate constant k0 increased linearly with the concentration of NaY, while it decreased drastically with the concentration of NaX. Some plausible mechanisms were examined, and the following mechanism was proposed. [Note to reader: Please see article pdf to view this scheme.] © 1998 John Wiley & Sons, Inc. Int J Chem Kinet 30: 523–532, 1998  相似文献   

3.
Methyl glyoxylate has been polymerized in CH2Cl2 solution either cationically or anionically to give the corresponding polyacetal. Thermodynamic and kinetic results of the polymerizations initiated by CF3SO3H, BF3OEt2, and NEt3 are reported here. Monomer conversion was followed by UV and active centers concentrations were determined through phosphorus end-capping. A ceiling temperature of 26°C was observed for Meq=l mol.H (Tc = 109°C for bulk) with ΔH°=−29.5 kJ.mo−1 and ΔS°=−99 J.mo−1.K−1. Both initiations were found instantaneous and quantitative, and no termination was observable within the time scale of the reaction (second to hour). Cationic propagation appears to take place essentially on free ions (kp+ ≈ 60 l.moH.s−1 at −20°C) in equilibrium with a larger amount (Kd ≈ 6.7.10−6 mol.−1) of much less reactive ion-pairs (k ≈ 0.16 ± 0.03 1.mo−1.s−1). A stopped-flow device was used to follow the much more rapid anionic polymerization (kp,app ≈ 8 ± 2.103 1.mo−1.s−1 at −20°C).  相似文献   

4.
The interaction of H+-ATPase complex ECF1·F0 and the Trk system of K+ accumulation was studied in E.coli grown quasi-anaerobically in peptone media with glucose (anaerobia) and aerobically in salt medium with succinate (aerobia). In anaerobia the Trk system takes part in H+-K+ exchange, displaying km = 3.7 mM, νmax =1.6 mM g−1 min−1 and in aerobia in the Trk system has Km = 3.4 mM, νmax = 0.45 mM g−1 min−1. The K+ accumulation is blocked by DCC in anaerobia and by cyanide together with DCC in aerobia, whereas protonophores and arsenate block the K+ uptake in bacteria grown under either condition. Valinomycin decreases the K+ accumulation in anaerobia and increases (or has no effect) that in aerobia. ECF1·F0 is sensitive and the Trk system is insensitive to variation of the external osmotic pressure in both cases. The ratio H+ : K+ is stable and equal to 2:1 in anaerobia and is changed from 0.5 to 5.0 in aerobia in response to variation of pH, K+ activity and temperature, Q10 is about 2.8 both for ECF1·F0 and for the Trk system in anaerobia, but 2.4 and near 1.0 respectively in aerobia. The distribution of K+ in anaerobia is 2500 (potassium equilibrium potential of − 210 mV) which is much more than the measured Um of −145 mV. The distribution of K+ in aerobia is 720, which is in good conformity with the measured membrane potential of −175 mV. The structural association of ECF1·F0 and the Trk system, forming a H+-K+ pump, has been assumed previously to take place in anaerobically grown E.coli; these systems operate separately in aerobic cells. According to Bakker and Helmer et al. the Trk system transports K+ along the electrical field, its function being regulated by ATP.  相似文献   

5.
Pseudo‐first‐order rate constants (kobs) for pH‐independent hydrolysis of phthalimide ( 1 ), obtained at a constant total concentration of cetyltrimethylammonium bromide and hydroxide ([CTABr]T), 2.0 × 10?4 M 1 , 0.02 M MOH (M+ = Li+, Na+ and K+) and various concentrations of inert salt MX (= LiCl, LiBr, NaCl, NaBr, KCl and KBr), follow a relationship derived from the pseudophase micellar (PM) model coupled with an empirical equation. This relationship gives empirical constants, FX/S and KX /S, with S representing anionic 1 . The magnitude of FX/S is the measure of the fraction of micellized anionic 1 (S?M) transferred to the aqueous phase by the limiting concentration of X?. The value of KX/S is the measure of the ability of the counterions (X?) to expel the reactive counterions (S?) from the cationic micellar surface to the aqueous phase. The values of FX/ S are ~ 1 for MBr (M+ = Li+, Na+ and K+) and in the range ? 0.7 to ? 0.5 for MCl (M+ = Na+ and K+) at 0.006, 0.010 and 0.016 M CTABr. For LiCl, the values of FX/S become ~1 at 0.006 and 0.010 M CTABr and 0.8 at 0.016 M CTABr. The values of the empirical constants, FX/S and KX/S, have been used to determine the usual ion exchange constant (KClBr). The mean values of KClBr are 3.9 ± 0.5, 2.7 ± 0.1, and 2.6 ± 0.3 for LiX, NaX, and KX, respectively. These values of KClBr are comparable with those obtained directly by other physicochemical techniques. Thus, this new method for the determination of ion exchange constants for various counterions of cationic micelles may be considered as a reliable one. © 2010 Wiley Periodicals, Inc. Int J Chem Kinet 43: 9–20, 2011  相似文献   

6.
The cation-insertion solid state electrochemistry of a potassium copper(II) hexacyanoferrate in contact with LiClO4/DMSO, NaPF6/DMSO, and KPF6/DMSO electrolytes has been theoretically and experimentally studied using the voltammetry of immobilized particles methodology. Voltammetric data, combined with SEM/EDS analysis permit to determine a K0.876CuII1.328FeIII0.049[FeIII0.318FeII0.682(CN)6] stoichiometry for the synthesized solid. Separation of electronic and ionic contributions to Gibbs energy changes can be made based on cyclic voltammetric and open circuit potential measurements. These parameters can be combined to measure values of the Gibbs energy of cation-independent electron transfer of 7.2±0.4 (K+), 7.1±0.5 (Na+) kJ mol−1, in close agreement with the expected independence of this parameter on the electrolyte cation. The reduction Fe(III) centers bound to cyano groups exhibit a cation-dependent, essentially Nernstian character which can be described in terms of Na+ and K+ insertion/deinsertion while in the case of Li+ electrolytes there is significant co-cation diffusion. Chronoamperometric data provide estimates of the diffusion coefficients of Na+, and K+ ions through the solid around 10−9 cm2 s−1.  相似文献   

7.
The present analysis was done to elucidate the electrical properties of an established mammalian epithelial cell line, i.e. the Madin Darby canine kidney (MDCK) cell. Rapid changes of the extracellular fluid were made to identify conductive pathways across the cell membrane. During control conditions mimicking in vivo extracellular fluid, the potential difference across the cell membrane (Um) approached −51.0±0.5 mV (n = 122). With potassium-sensitive muelectrodes, a potential difference of +14,4 ± 1.4 mV (n = 5) was determined. Rapid increase of the extracellular potassium concentration from 5.4 to 20 mmol/dm3 depolarized Um (dUmK) by +17.0 ± 0.4 mV (n = 85). Accordingly, some 50% of the cell membrane conductance is due to potassium conductive pathways, i.e. the transference number for potassium (tK) approaches 0.5. When 10 μmol/dm3 valinomycin was added to the extracellular fluid, Um increased gradually to −74.7 ± 0.8 mV and tK increased to 0.86. Barium at a concentration of 1 mmol/dm3 in the extracellular fluid depolarized Um by +20.2 ± 0.4 mV and virtually abolished tK. However, when extracellular potassium was increased from 5.4 to 35 mmol/dm3, further depolarization was observed, indicating that barium can be displaced by high potassium concentrations. Rapid reduction of the extracellular sodium concentration from 133 to 19 mmol/dm3 led to a significant sustained hyperpolarization (dUmNa) of −13.5±0.6 mV (n =15). Additional experiments indicated that the hyperpolarization was at least partly due to opening of potassium conductive pathways.In conclusion, the cell membrane conductance of MDCK cells is partially due to a potassium conductance which can be blocked by barium and stimulated by the reduction of extracellular sodium. Intracellular potassium is higher than that for electrochemical equilibrium across the cell membrane, and thus potassium diffusion is directed from the cell to the extracellular fluid.  相似文献   

8.
Affinity capillary electrophoretic (ACE) study has proved the selectivity of hexaarylbenzene-based polyaromatic receptor (R) for K+ ion over Na+ ion. The apparent binding constants of the R complexes with K+ and Na+ ions were determined from the dependence of effective electrophoretic mobility of R on the concentration of the above alkali metal ions in the background electrolyte using a non-linear regression analysis. The apparent binding constants (Kb) of the K-R+ and Na–R+ complexes in methanolic medium were evaluated as log Kb = 3.20 ± 0.22 for the K–R+ complex, and log Kb??0.7 for the Na–R+ complex.  相似文献   

9.
《Tetrahedron letters》1997,38(28):5021-5024
1,4,7,10-Tetraazacyclododecane derivatives having ester- and amide-functionalized sidearms predominantly extracted Na+ ion from an aqueous mixture of Li+, Na+ and K+ ions into a CH2Cl2 via specific octadentate complexation.  相似文献   

10.
The kinetic of D,L-lactide polymerization in presence of biocompatible zirconium acetylacetonate initiator was studied by differential scanning calorimetry in isothermal mode at various temperatures and initiator concentrations. The enthalpy of D,L-lactide polymerization measured directly in DSC cell was found to be ΔH=−17.8±1.4 kJ mol−1. Kinetic curves of D,L-lactide polymerization and propagation rate constants were determined for polymerization with zirconium acetylacetonate at concentrations of 250–1000 ppm and temperature of 160–220 °C. Using model or reversible polymerization the following kinetic and thermodynamic parameters were calculated: activation energy Ea=44.51±5.35 kJ mol−1, preexponential constant lnA=15.47±1.38, entropy of polymerization ΔS=−25.14 J mol−1 K−1. The effect of reaction conditions on the molecular weight of poly(D,L-lactide) was shown.  相似文献   

11.
The theoretical calculations have predicted that nonmetal‐doped potassium clusters can be used in the synthesis of a new class of charge‐transfer salts which can be considered as potential building blocks for the assembly of novel nanostructured material. In this work, KnCl (n = 2–6) and KnCln?1 (n = 3 and 4) clusters were produced by vaporization of a solid potassium chloride salt in a thermal ionization mass spectrometry. The ionization energies (IEs) were measured, and found to be 3.64 ± 0.20 eV for K2Cl, 3.67 ± 0.20 eV for K3Cl, 3.62 ± 0.20 eV for K4Cl, 3.57 ± 0.20 eV for K5Cl, 3.69 ± 0.20 eV for K6Cl, 3.71 ± 0.20 eV for K3Cl2 and 3.72 ± 0.20 eV for K4Cl3. The KnCl+ (n = 3–6) clusters were detected for the first time in a cluster beam generated by the thermal ionization source of modified design. Also, this work is the first to report experimentally obtained values of IEs for KnCl+ (n = 3–6) and KnCln?1+ (n = 3 and 4) clusters. The ionization energies for KnCl+ and KnCln?1+ clusters are much lower than the 4.34 eV of the potassium atom; hence, these clusters should be classified as ‘superalkali’ species. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

12.
An electrochemical cell of potentiometric type Na0.5WO3 (reference electrode)/Na+-solid electrolyte/PbS (working electrode) capable of rapid and selective changing of the electromotive force value owing to H2S concentration variations in gas surroundings has been investigated at 295±1 K and a relative humidity of 52%. The sensitivity of this cell was 130 mV/decade at a H2S concentration within the range 13–130 ppm. Sodium-conducting solid electrolytes of Na3Zr2Si2PO12 and Na5GdSi4O12 compositions were used as the Na+ solid electrolyte. Such a cell can be used for analysis of H2S containing water solutions when the reference electrode and the Na+ solid electrolyte are thoroughly isolated from the surroundings. Electronic Publication  相似文献   

13.
Unlike many other biologically relevant ions (Na+, K+, Ca2+, Cl, etc) and protons, whose cellular concentrations are closely regulated by highly selective channel proteins, Li+ ion is unusual in that its concentration is well tolerated over many orders of magnitude and that no lithium-specific channel proteins have so far been identified. While one naturally evolved primary pathway for Li+ ions to traverse across the cell membrane is through sodium channels by competing with Na+ ions, highly sought-after artificial lithium-transporting channels remain a major challenge to develop. Here we show that sulfur-containing organic nanotubes derived from intramolecularly H-bonded helically folded aromatic foldamers of 3.6 Å in hollow cavity diameter could facilitate highly selective and efficient transmembrane transport of Li+ ions, with high transport selectivity factors of 15.3 and 19.9 over Na+ and K+ ions, respectively.  相似文献   

14.
In the last 30 years, fullerene-based materials have become popular building blocks for devices with a broad range of applications. Among fullerene derivatives, endohedral metallofullerenes (EMFs, M@Cx) have been widely studied owing to their unique properties and reactivity. For real applications, fullerenes and EMFs must be exohedrally functionalized. It has been shown that encapsulated metal cations facilitate the Diels–Alder reaction in fullerenes. Herein, the Bingel–Hirsch (BH) addition of ethyl bromomalonate over a series of ion-encapsulated M@C60 (M=Ø, Li+, Na+, K+, Mg2+, Ca2+, and Cl; Ø@C60 stands for C60 without any endohedral metal) is quantum mechanically explored to analyze the effect of these ions on the BH addition. The results show that the incarcerated ion has a very important effect on the kinetics and thermodynamics of this reaction. Among the systems studied, K+@C60 is the one that leads to the fastest BH reaction, whereas the slowest reaction is given by Cl@C60.  相似文献   

15.
Standard values of Gibbs free energy, entropy, and enthalpy of Na2Ti6O13 and Na2Ti3O7 were determined by evaluating emf-measurements of thermodynamically defined solid state electrochemical cells based on a Na–β″-alumina electrolyte. The central part of the anodic half cell consisted of Na2CO3, while two appropriate coexisting phases of the ternary system Na–Ti–O are used as cathodic materials. The cell was placed in an atmosphere containing CO2 and O2. By combining the results of emf-measurements in the temperature range of 573⩽T/K⩽1023 and of adiabatic calorimetric measurements of the heat capacities in the low-temperature region 15⩽T/K⩽300, the thermodynamic data were determined for a wide temperature range of 15⩽T/K⩽1100. The standard molar enthalpy of formation and standard molar entropy at T=298.15 K as determined by emf-measurements are ΔfHm0=(−6277.9±6.5) kJ · mol−1 and Sm0=(404.6±5.3) J · mol−1 · K−1 for Na2Ti6O13 and ΔfHm0=(−3459.2±3.8) kJ · mol−1 and Sm0=(227.8±3.7) J · mol−1 · K−1 for Na2Ti3O7. The standard molar entropy at T=298.15 K obtained from low-temperature calorimetry is Sm0=399.7 J · mol−1 · K−1 and Sm0=229.4 J · mol−1 · K−1 for Na2Ti6O13 and Na2Ti3O7, respectively. The phase widths with respect to Na2O content were studied by using a Na2O-titration technique.  相似文献   

16.
A highly K+‐selective two‐photon fluorescent probe for the in vitro monitoring of physiological K+ levels in the range of 1–100 mM is reported. The two‐photon excited fluorescence (TPEF) probe shows a fluorescence enhancement (FE) by a factor of about three in the presence of 160 mM K+, independently of one‐photon (OP, 430 nm) or two‐photon (TP, 860 nm) excitation and comparable K+‐induced FEs in the presence of competitive Na+ ions. The estimated dissociation constant (Kd) values in Na+‐free solutions (KdOP=(28±5) mM and KdTP=(36±6) mM ) and in combined K+/Na+ solutions (KdOP=(38±8) mM and KdTP=(46±25) mM ) reflecting the high K+/Na+ selectivity of the fluorescent probe. The TP absorption cross‐section (σ2PA) of the TPEF probe+160 mM K+ is 26 GM at 860 nm. Therefore, the TPEF probe is a suitable tool for the in vitro determination of K+.  相似文献   

17.
《Electroanalysis》2004,16(21):1785-1790
Binaphthyl‐based crown ethers incorporating anthraquinone, benzoquinone, and 1,4‐dimethoxybezene have been synthesized and tested for Rb+ selective ionophores in the poly(vinyl chloride) (PVC) membrane. The membrane containing NPOE gave a better Rb+ selectivity than those containing either DOA or BPPA as a plasticizer. The response was linear within the concentration range of 1.0×10?5–1.0×10?1 M and the slope was 54.7±0.5 mV/dec. The detection limit was determined to be 9.0×10?6 M and the optimum pH range of the membrane was 6.0–9.0. The ISE membrane exhibits good selectivity for Rb+ over ammonium, alkali metal, and alkaline earth metal ions. Selectivity coefficients for the other metal ions, log KPot were ?2.5 for Li+, ?2.4 for Na+, ?2.0 for H+, ?1.0 for K+, ?1.2 for Cs+, ?1.6 for NH4+, ?4.5 for Mg2+, ?5.0 for Ca2+,?4.9 for Ba2+. The lifetime of the membrane was about one month.  相似文献   

18.
The far-IR spectrum from 375 to 30 cm−1 of gaseous 3-chloro-2-methylpropene, CH2=C(CH3)CH2Cl, has been recorded at a resolution of 0.10 cm−1. The fundamental asymmetric torsional mode for the gauche conformer is observed at 84.3 cm−1 with three excited states falling to lower frequency. For the higher energy s-cis conformer, where the chlorine atom eclipses the double bond, the asymmetric torsion is observed at 81.3 cm−1 with two excited states falling to lower frequency. Utilizing the s-cis and gauche torsional frequencies, the gauche dihedral angle and the enthalpy difference between conformers, the potential function governing the interconversion of the rotamers has been calculated. The determined potential function coefficients are (in reciprocal centimeters): V1=189±12, V2=−358±11, V3=886±2 and V4=−12±2 with an enthalpy difference between the more stable gauche and s-cis conformers of 150 ±25 cm−1 (430 ± 71 cal mol−1). This function gives values of 661 cm−1 (1.89 kcal mol−1), 1226 cm−1 (3.51 kcal mol−1) and 812 cm−1 (2.32 kcal mol−1), for the s-cis to gauche, gauche to gauche, and gauche to s-cis barriers, respectively. From the methyl torsional frequency of 170 cm−1 for the gauche conformer, the threefold barrier of 678 cm−1 (1.94 kcal mol−1) has been calculated. The asymmetric potential function, conformational energy difference and optimized geometries of both conformers have also been obtained from ab initio calculations with both the 3–21G* and 6–31G* basis sets. A normal-coordinate analysis has also been performed with a force field determined from the 3–21G* basis set. These data are compared with the corresponding data for some similar molecules.  相似文献   

19.
The six possible heats of mixing of the system Na+, Pr4N+?Br?, I? have been measured at 35°C in the solvent N-methylacetamide at an ionic strength of 0.5. The cross-square rule of T. F. Young(1,2) does not hold accurately, and there are strong attractive forces between a cation and an anion. The largest effect in the heats of mixing is the formation of tetra-n-propylammonium iodide interactions. The data is compared to that of the Na+,K+?Cl?,NO 3 ? system in water.  相似文献   

20.
To obtain reliable thermodynamic data for Na2S(s), solid-state EMF measurements of the cell Pd(s)|O2(g)|Na2S(s), Na2SO4(s)|YSZ| Fe(s), FeO(s)|O2(g)ref| Pd(s) were carried out in the temperature range 870 < T/K < 1000 with yttria stabilized zirconia as the solid electrolyte. The measured EMF values were fitted according to the equation Efit/V (±0.00047) = 0.63650  0.00584732(T/K) + 0.00073190(T/K) ln (T/K). From the experimental results and the available literature data on Na2SO4(s), the equilibrium constant of formation for Na2S(s) was determined to be lg Kf(Na2S(s)) (±0.05) = 216.28  4750(T/K)−1  28.28878 ln (T/K). Gibbs energy of formation for Na2S(s) was obtained as ΔfG(Na2S(s))/(kJ · mol−1) (±1.0) = 90.9  4.1407(T/K) + 0.5415849(T/K) ln (T/K). By applying third law analysis of the experimental data, the standard enthalpy of formation of Na2S(s) was evaluated to be ΔfH(Na2S(s), 298.15 K)/(kJ · mol−1) (±1.0) = −369.0. Using the literature data for Cp and the calculated ΔfH, the standard entropy was evaluated to S(Na2S(s), 298.15 K)/(J · mol−1 · K−1) (±2.0) = 97.0.  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