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 562 毫秒
1.
Influence of the Composition of Mixed Solvents on the Stability and Formation Constants of Copper (II) and Nickel(II) Complexes of Substituted 1,2-Dioximes The stability constants cK1, cK2, and cβ2 of the complexes which are formed in the systems M2+/DH2, M2+/Ac? DH2, and M2+/Et, Me\documentclass{article}\pagestyle{empty}\begin{document}$ \mathop {\rm N}\limits^ \oplus $\end{document}? DH2 (M2+ ? Cu2+, Ni2+; DH2, Ac? DH2, Et2Me\documentclass{article}\pagestyle{empty}\begin{document}$ \mathop {\rm N}\limits^ \oplus $\end{document}? DH2 = 1,2-dioximes) are determined in water and in water-dioxane mixtures (25, 50 and 75 per cent). Because of the stabilisation of the l,2-complexes by intramolecular hydrogen bonds cK2, is always higher than cK1. On account of the decrease of the dielectricity constant the constants cK1, cK2, and cβ2, rise with increasing contents of dioxane in the mixtures. The influence of the dielectricity constant may be eliminated by considering the formation constants cK1(B), cK2(B), and cβ2(B). The individual formation constants cK1(B) of the 1,l-complexes investigated are independent of the composition of the solvent, but among the overall formation constants cβ2(B) this comes true only for the complexes Ni(Ac? DH)2, Ni(Et2Me\documentclass{article}\pagestyle{empty}\begin{document}$ \mathop {\rm N}\limits^ \oplus $\end{document}? DH)2, and Cu(Et2Me\documentclass{article}\pagestyle{empty}\begin{document}$ \mathop {\rm N}\limits^ \oplus $\end{document}? DH)2. With Cu(DH2) and Cu(Ac? DH)2 a linear relation between log cβ2(B) and the molar fraction of water is stated. This effect is attributed to a specific solvatation of the chelates Cu(DH)2 and Cu(Ac? DH)2 by water.  相似文献   

2.
Precipitation or coprecipitation of polyelectrolytes has been largely investigated. However, the precipitation of polyelectrolytes via addition of charged and non‐charged surfactants has not been systematically studied and reported. Consequently, the aim of this work is to investigate the effect of different surfactants (anionic, cationic, non‐charged and zwitterionic) on the precipitation of cationic and anionic polymethylmethacrylate polymers (Eudragit). The surfactants effect has been investigated as a function of their concentration. Special attention has been dedicated to the CMC range and to the colloidal characterization of the formed dispersions. Moreover, the effect of salt (NaCl) and pH was also addressed. It is pointed out that non‐ionic and zwitterionic surfactants do not interact with charged Eudragit E100 and L100. For oppositely charged Eudragit E100/SDS and Eudragit L100/CTAB, precipitation occurs, and the obtained dispersions have been characterized in terms of particle size distribution and zeta potential. It was established that the binding of SDS molecules to Eudragit E100 polymer chains is made through the negative charges of the surfactant heads under the CMC value whereas binding of CTAB to Eudragit L100 chains is made at a CTAB concentration 5 times above its CMC. For Eudragit E100/SDS system, a more acidic medium induces aggregation. A same result was observed for the Eudragit L100/CTAB at a more basic pH. Moreover, it was observed that increasing salt concentration (higher than 100 mM) led to aggregation as generally observed for polycations/anionic surfactant systems.  相似文献   

3.
The Debye–Hückel and non-Debye–Hückel contributions to the Gibbs energy interaction parameters are investigated for electrolyte (E) + non-electrolyte (N) + water (W) systems. A method is proposed for calculating the interaction parameters, C n DH,C n N, and C n T, which represent the Debye–Hückel, non-Debye–Hückel, and total contributions, respectively. Four ternary E + N + W systems are chosen and the interaction parameters are computed with different forms of the Debye–Hückel equation. Results show that: (1) the Gibbs energy interaction parameters between E and N can be divided into two parts: the Debye–Hückel contribution and the non-Debye–Hückel contribution, C n T=C n DH+C n N; (2) the signs and magnitudes of the Debye–Hückel contribution to the interaction parameters, C n DH, depend mainly on the change in the dielectric constant of the solvent due to the addition of the non-electrolyte into the solvent; and (3) when the addition of the non-electrolyte only affects slightly the dielectric constant of the solvent, C 1DH (indicating the Debye–Hückel contribution to the interaction parameter for E + N) has a very small value and consequently can be neglected. In general, C 1DH is large, even larger than C 1N. Electronic Supplementary Material The online version of this article () contains supplementary material, which is available to authorized users.  相似文献   

4.
Osmotic vapor pressure measurements have been carried out for three ternary systems, H2O + 0.2 m 18-crown-6 + LiCl, H2O + 0.2 m 18-crown-6 + NaCl and H2O + 0.2 m 18-crown-6 + KCl at 298.15 K using vapor pressure osmometry. Water activities for each ternary system were measured and used to calculate the activity coefficients of 18-crown-6 (18C6) and its salts following the methodology developed by Robinson and Stokes for isopiestic measurements. In the concentration range studied, it was found that (in NaCl and KCl solutions) there is considerable lowering of activity coefficients of one component in the presence of other solutes that has been attributed to the formation of the complexed 18C6:Na+ (or 18C6:K+) species in solution. The Gibbs energies of transfer of alkali chlorides from water to aqueous 18C6 solutions and that of 18C6 from water to aqueous electrolyte solutions have been calculated. These were further used to evaluate the pair and triplet interaction parameters. The calculation of thermodynamic equilibrium constants using the pair interaction parameter, g NE (i.e., the nonelectrolyte–electrolyte pair interaction) for the studied complexation of cations yields values which are in good agreement with those reported in literature obtained by using ion-selective potentiometry and calorimetry. The results are discussed in terms of water structural effects, complex formation, and hydrophobic interactions.  相似文献   

5.
Photodynamic therapy combines visible light and a photosensitizer (PS) in the presence of molecular oxygen to generate reactive oxygen species able to modify biological structures such as phospholipids. Phosphatidylethanolamines (PEs), being major phospholipid constituents of mammalian cells and membranes of Gram‐negative bacteria, are potential targets of photosensitization. In this work, the oxidative modifications induced by white light in combination with cationic porphyrins (Tri‐Py+‐Me‐PF and Tetra‐Py+‐Me) were evaluated on PE standards. Electrospray ionization mass spectrometry (ESI‐MS) and tandem mass spectrometry (ESI‐MS/MS) were used to identify and characterize the oxidative modifications induced in PEs (POPE: PE 16:0/18:1, PLPE: PE 16:0/18:2, PAPE: PE 16:0/20:4). Photo‐oxidation products of POPE, PLPE and PAPE as hydroxy, hydroperoxy and keteno derivatives and products due to oxidation in ethanolamine polar head were identified. Hydroperoxy‐PEs were found to be the major photo‐oxidation products. Quantification of hydroperoxides (PE‐OOH) allowed differentiating the potential effect in photodamage of the two porphyrins. The highest amounts of PE‐OOH were notorious in the presence of Tri‐Py+‐Me‐PF, a highly efficient PS against bacteria. The identification of these modifications in PEs is an important key point in the understanding cell damage processes underlying photodynamic therapy approaches. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

6.
The hydrolysis of (C2H5)2Sn2+, (C2H5)3Sn+ and (n‐C3H7)3Sn+ has been studied, by potentiometric measurements ([H+]‐glass electrode), in NaNO3, NaCl, NaCl/Na2SO4 mixtures and in a synthetic seawater (SSWE), as an ionic medium simulating the major composition of natural seawater, at different ionic strengths (0 ≤ I ≤ 5 mol dm?3) and salinities (15 ≤ S ≤ 45), and at t = 25 °C. Five hydrolytic species for (C2H5)2Sn2+, three for (C2H5)3Sn+ and two for (C3H7)3Sn+ are found. Interactions with the anion components of SSWE, considered as single‐salt seawater, are determined by means of a complex formation model. A predictive equation for the calculation of unknown hydrolysis constants of trialkyltin(IV) cations, such as tributyltin(IV), in NaNO3, NaCl, and SSWE media at different ionic strengths is proposed. Equilibrium constants obtained are also used to determine the interaction parameters of Pitzer equations. Copyright © 2001 John Wiley & Sons, Ltd.  相似文献   

7.
A density functional theory computational chemistry study has revealed a fundamental structural difference between [Ti(Cp)3]+ and its congeners [Zr(Cp)3]+ and [Hf(Cp)3]+/(Cp=cyclopentadienyl). Whereas the latter two are found to contain three uniformely η5-coordinated Cp ligands (3η5-structural type), [Ti(Cp)3]+ is shown to prefer a 2η5η2 structure. [Ti(Cp)3]+[B(C6F5)3(Me)] ( 10 ⋅[B(C6F5)3(Me)]) was experimentally generated by treatment of [Ti(Cp)3(Me)] ( 7a ) with B(C6F5)3 (Scheme 3). Low-temperature 1H-NMR spectroscopy in CDFCl2 (143 K, 600 MHz; Fig. 8) showed a splitting of the Cp resonance into five lines in a 2 : 5 : 2 : 5 : 1 ratio which would be in accord with the theoretically predicted 2η5η2-type structure of [Ti(Cp)3]+. The precursor [Ti(Cp)3(Me)] ( 7a ) exhibits two 1H-NMR Cp resonances in a 10 : 5 ratio in CD2Cl2 at 223 K. Treatment of [HfCl(Cp)2(Me)] ( 6c ) with sodium cyclopentadienide gave [Hf(Cp)3(Me)] ( 7c ) (Scheme 1). Its reaction with B(C6F5)3 furnished the salt [Hf(Cp)3]+[B(C6F5)3(Me)] ( 8 ⋅[B(C6F5)3(Me)]), which reacted with tert-butyl isocyanide to give the cationic complex [Hf(Cp)3(C=N−CMe3)]+ ( 9a ; with counterion [B(C6F5)3(Me)] (Scheme 2). Complex cation 9a was characterized by X-ray diffraction (Fig. 7). Its Hf(Cp3) moiety is of the 3η5-type. The structure is distorted trigonal-pyramidal with an average D−Hf−D angle of 118.8° and an average D−Hf−C(1) angle of 96.5° (D denotes the centroids of the Cp rings; Table 6). Cation 9a is a typical d0-isocyanide complex exhibiting structural parameters of the C≡N−CMe3 group (d(C(1)−N(2))=1.146 (5) Å; IR: v˜(C≡N) 2211 cm−1) very similar to free uncomplexed isonitrile. Analogous treatment of 8 with carbon monoxide yielded the carbonyl (d0-group-4-metal) complex [Hf(Cp)3(CO)]+ ( 9b ; with counterion [B(C6F5)3(Me)]) (Scheme 2) that was also characterized by X-ray crystal-structure analysis (Fig. 6). Complex 9b is also of the 3η5-structural type, similar to the peviously described cationic complex [Zr(Cp)3(CO)]+, and exhibits properties of the CO ligand (d(C−O)=1.11 (2) Å; IR: v˜(C≡O) 2137 cm−1) very similar to the free carbon monoxide molecule.  相似文献   

8.
Ruthenium(II) bisbipyridyl complexes cis-[Ru(bpy)2(L)NO2](BF4) (bpy is 2,2'-bipyridyl) with 4-substituted pyridine ligands L = 4-(Y)py (Y = NH2, Me, Ph, and CN) were obtained. The equilibrium constants of the reversible nitro-nitrosyl transition [Ru(bpy)2(L)NO2]+ + 2H+ [Ru(bpy)2(L)NO]3 + + H2O were measured in solutions with pH 1.5-8.5 (ionic strength 0.4). The constants correlate with the protonation constants of free ligands 4-(Y)py.  相似文献   

9.
The weak association between sodium and carbonate ions has been investigated at 25°C using high-precision sodium ion-selective electrode potentiometry in solutions of ionic strength ranging from 0.5 to 7.0 M in CsCl and in 1.0 M Me4NCl media. The protonation constants of CO 3 2- (aq) were also measured, using a H+-responsive glass electrode in 1.0 M Me4NCl and NaCl. The value of the ion-pair association constant calculated from the difference in the protonation constants in these two media was in excellent agreement with that obtained from the Na+ISE measurements. Evidence is also presented for the formation of extremely weak ion pairs between Na+ and HCO 3 - and between Cs+ and CO 3 2- .  相似文献   

10.
Stability constants of macromolecular metallocomplexes of transition metal ions (Ag+, Cu2+, Ni2+, Fe3+) with sulfonated polymers in water and aqueous HCl and NaCl solutions were determined from quenching by transition metal ions of the luminescence of macromolecules labeled with luminescent groups.  相似文献   

11.
The use of lithium cation in composites of block copolymers [polyethylene‐b‐polyethylene oxide (PE‐b‐50%PEO and PE‐b‐80%PEO)] and their derivatives was tested as a modifier of the vapor sorption and impedance of these complexes. The block copolymer PE‐b‐80%PEO was modified by oxidation of its hydroxyl end group to both a carboxylic acid group (PE‐b‐80%PEO)CH2COOH and its sodium salt (PE‐b‐80%PEO)CH2COO? Na+ for the purpose of improving its compatibility and performance as a matrix for composites. These modified copolymers were characterized by FTIR, DSC, and mass spectrometry. The sorption of water of these copolymers and their composites with lithium nitrate was also compared, as well as the electrical properties of their composites were measured by electrical impedance spectroscopy. For the composites obtained with PE‐b‐80%PEO and lithium nitrate, it was found that lithium cation plays an important role increasing the sorption rate, which is maximized for the PE‐b‐80%PEO + (21% lithium nitrate) composite. For the copolymers (PE‐b‐80%PEO)CH2COOH and (PE‐b‐80%PEO)CH2COO? Na+ and their composites, the highest sorption rate was observed for salt in the following order: COO? Na+ > COOH > OH. The PE‐b‐80%PEO + (21% lithium nitrate) composite behaves as a solid polymeric ionic conductor fitting the Williams–Landel–Ferry equation. However, both (PE‐b‐80% PEO)CH2COOH and (PE‐b‐80%PEO)CH2COO? Na+ + (21% lithium nitrate) composites fitted the Variable Range Hopping equation, indicating a conductance trend with temperature governed by a thermally activated with energy of 0.482 and 0.524 eV and not by a relaxation process. © 2010 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 48: 1809–1817, 2010  相似文献   

12.
The effect of electrolytes (NaCl and CaCl2) and polymers (CPAM and HPAM) on the thixotropy of Mg‐Al‐layered double hydroxide (LDHs)/kaolinite dispersions has been investigated. It was observed that the type of thixotropy in LDH/kaolinite dispersions may be affected by NaCl, but not by CaCl2 in range of concentration of interest. The type of thixotropy in LDH/kaolinite dispersion with R=0 transformed from positive thixotropy to complex thixotropy and at last positive thixotropy again with the concentration of NaCl in range of 0.00–0.10 mol·L−1; the type of thixotropy in LDHs/kaolinite dispersions with R=0.25 transformed from complex thixotropy to positive thixotropy and then complex thixotropy again with the concentration of NaCl in range of 0.00–0.10 mol·L−1. The type of thixotropy in LDH/kaolinite dispersion with R=0 may be not affected by cationic polyacrylamide (CPAM) and hydrolyzed polyacrylamide (HPAM); but the LDHs/kaolinite dispersions with R=0.25 transformed from complex thixotropy to positive thixotropy with the both polymers concentration in range of interest, which indicated that the microstructure of the dispersion changed from weak folc sediments structure to steric network structure.  相似文献   

13.
A three-dimensional potential energy function has been calculated for the X1Σ+g state of NO+2 from ab initio MRD-CI data. With this PE function, converged vibrational calculations have also been performed for ten vibrational states, with the aid of a computer program developed in the present work for this purpose. The calculated harmonic frequencies, vibrational term values and rotational constants are in good agreement with experimental data.  相似文献   

14.
New silver(I) acylpyrazolonato derivatives displaying a mononuclear, polynuclear, or ionic nature, as a function of the ancillary azole ligands used in the synthesis, have been fully characterized by thermal analysis, solution NMR spectroscopy, solid‐state IR and NMR spectroscopies, and X‐ray diffraction techniques. These derivatives have been embedded in polyethylene (PE) matrix, and the antimicrobial activity of the composite materials has been tested against three bacterial strains (E. coli, P. aeruginosa, and S. aureus): Most of the composites show antimicrobial action comparable to PE embedded with AgNO3. Tests by contact and release tests for specific migration of silver from PE composites clearly indicate that, at least in the case of the PE, for composites containing polynuclear silver(I) additives, the antimicrobial action is exerted by contact, without release of silver ions. Moreover, PE composites can be re‐used several times, displaying the same antimicrobial activity. Membrane permeabilization studies and induced reactive oxygen species (ROS) generation tests confirm the disorganization of bacterial cell membranes. The cytotoxic effect, evaluated in CD34+ cells by MTT (3‐(4,5‐dimethylthiazole‐2‐yl)‐2,5‐diphenyltetrazoliumbromide) and CFU (colony forming units) assays, indicates that the PE composites do not induce cytotoxicity in human cells. Studies of ecotoxicity, based on the test of Daphnia magna, confirm tolerability of the PE composites by higher organisms and exclude the release of Ag+ ions in sufficient amounts to affect water environment.  相似文献   

15.
From extraction experiments and γ-activity measurements, the exchange extraction constants corresponding to the general equilibrium M+(aq)+NaL+(nb)⇔ML+(nb)+Na+(aq) taking place in the two-phase water-nitrobenzene system [M+=Li+, K+, Rb+, Cs+; L = p-tert-butylcalix[4]arene-tetrakis (N, N-dimethylthioacetamide); aq = aqueous phase, nb = nitrobenzene phase] were evaluated. Furthermore, the stability constants of the ML+ complexes in water saturated nitrobenzene were calculated; they were found to increase in the cation order Cs+<Rb+<K+<Li+<Na+.  相似文献   

16.
The gas‐phase dehydration–rearrangement (DR) reactions of protonated alcohols [Me2(R)CCH(OH2)Me]+ [R=Me ( ME ), Et ( ET ), and iPr ( I‐PR )] were studied by using static approaches (intrinsic reaction coordinate (IRC), Rice–Ramsperger–Kassel–Marcus theory) and dynamics (quasiclassical trajectory) simulations at the B3LYP/6‐31G(d) level of theory. The concerted mechanism involves simultaneous water dissociation and alkyl migration, whereas in the stepwise reaction pathway the dehydration step leads to a secondary carbocation intermediate followed by alkyl migration. Internal rotation (IR) can change the relative position of the migrating alkyl group and the leaving group (water), so distinct products may be obtained: [Me(R)CCH(Me)Me ??? OH2]+ and [Me(Me)CCH(R)Me ??? OH2]+. The static approach predicts that these reactions are concerted, with the selectivity towards these different products determined by the proportion of the conformers of the initial protonated alcohols. These selectivities are explained by the DR processes being much faster than IR. These results are in direct contradiction with the dynamics simulations, which indicate a predominantly stepwise mechanism and selectivities that depend on the alkyl groups and dynamics effects. Indeed, despite the lifetimes of the secondary carbocations being short (<0.5 ps), IR can take place and thus provide a rich selectivity. These different selectivities, particularly for ET and I‐PR , are amenable to experimental observation and provide evidence for the minor role played by potential‐energy surface and the relevance of the dynamics effects (non‐IRC pathways, IR) in determining the reaction mechanisms and product distribution (selectivity).  相似文献   

17.
Stability constants ( 1 NB ) of the 1:1 cationic complexes of Li+ Na+, K+ Ca2+ Sr2+ and Ba2+ with benzo-18-crown-6 (B18C6), Ca2+ and Sr2+ with 18C6 and dibenzo-18C6 and Li+, Na+, Ca2+, Sr2+ and Ba2+ with dibenzo-24-crown-8 in a nitrobenzene (NB) solution saturated with water (w) were determined at 25°C by ion-transfer polarography. From these values, distribution constants (K D,ML) of the 18C6-derivative complex cations between the w- and NB-phases were evaluated using the thermodynamic relation:K D,ML =K 1 NB , whereK (mol dm–3) is an overall equilibrium constant of the processes related to the complexation in the w-phase. The data on the distribution of the 18C6-derivative complex cations between the two phases and the complexation in the NB-phase were examined on the basis of an increase in the number of water molecules hydrated to the species relevant to these processes. The 18C6 derivatives showed higher solubilities in the NB-phase than in the w-phase by complexing with the univalent-metal ions, while, for the divalent-metal ions, the derivatives showed lower solubilities in the NB-phase.  相似文献   

18.
The cationic cluster complexes [Ru3(μ‐H)(μ‐κ2N,C‐L1 Me)(CO)10]+ ( 1 +; HL1 Me=N‐methylpyrazinium), [Ru3(μ‐H)(μ‐κ2N,C‐L2 Me)(CO)10]+ ( 2 +; HL2 Me=N‐methylquinoxalinium), and [Ru3(μ‐H)(μ‐κ2N,C‐L3 Me)(CO)10]+ ( 3 +; HL3 Me=N‐methyl‐1,5‐naphthyridinium), which contain cationic N‐heterocyclic ligands, undergo one‐electron reduction processes to become short lived, ligand‐centered, trinuclear, radical species ( 1 – 3 ) that end in the formation of an intermolecular C? C bond between the ligands of two such radicals, thus leading to neutral hexanuclear derivatives. These dimerization processes are selective, in the sense that they only occur through the exo face of the bridging ligands of trinuclear enantiomers of the same configuration, as they only afford hexanuclear dimers with rac structures (C2 symmetry). The following are the dimeric products that have been isolated by using cobaltocene as reducing agent: [Ru6(μ‐H)26‐κ4N2,C2‐(L1 Me)2}(CO)18] ( 5 ; from 1 +), [Ru6(μ‐H)26‐κ4N2,C2‐(L2 Me)2}(CO)18] ( 6 ; from 2 +), and [Ru6(μ‐H)24‐κ8N2,C6‐(L3 Me)2}(CO)18] ( 7 ; from 3 +). The structures of the final hexanuclear products depend on the N‐heterocyclic ligand attached to the starting materials. Thus, although both trinuclear subunits of 5 and 6 are face‐capped by their bridging ligands, the coordination mode of the ligand of 5 is different from that of the ligand of 6 . The trinuclear subunits of 7 are edge‐bridged by its bridging ligand. In the presence of moisture, the reduction of 3 + with cobaltocene also affords a trinuclear derivative, [Ru3(μ‐H)(μ‐κ2N,C‐L3′ Me)(CO)10] ( 8 ), whose bridging ligand (L3′ Me) results from the formal substitution of an oxygen atom for the hydrogen atom (as a proton) that in 3 + is attached to the C6 carbon atom of its heterocyclic ligand. The results have been rationalized with the help of electrochemical measurements and DFT calculations, which have also shed light on the nature of the odd‐electron species, 1 – 3 , and on the regioselectivity of their dimerization processes. It seems that the sort of coupling reactions described herein requires cationic complexes with ligand‐based LUMOs.  相似文献   

19.
《Analytical letters》2012,45(7):753-763
Abstract

The formal potential of Ag [222] /Ag(s) system is supposed to be independent of the solvent, taking in account the [222] ligand structure and the corresponding cryptate. This extrathermodynamic hypothesis is confronted to the one usually considered, water, methanol, D. M. S. O., acetonitrile and tetramethylurea being the used solvents. It seems possible to generalize this hypothesis to other cations and cryptands. The stability constants of cryptate [222] of the evaluation of pK Ag[222]+ constants, Nernst's law being respected. The argentometric titration of ligand [222] in presence of Na+, Li+, Tl+, Cd2+, Ni2+ ions allows to evaluate the pK of corresponding cryptates in D. M. S. O. and methanol. Tl [222]+/Tl system follows Nernst's law in methanol and D. M. S. O. The dissociation constants have been evaluated from polarographic measurements in acetonitrile. The electrochemical systems are not rigorously fast, which does not allow an accurate determination of those constants by that method. Still it gives an order of magnitude.  相似文献   

20.
The facilitated transfer of alkali metal ions (Na+, K+, Rb+, and Cs+) by 25,26,27,28‐tetraethoxycarbonylmethoxy‐thiacalix[4]arene across the water/1,2‐dichloroethane interface was investigated by cyclic voltammetry. The dependence of the half‐wave transfer potential on the metal and ligand concentrations was used to formulate the stoichiometric ratio and to evaluate the association constants of the complexes formed between ionophore and metal ions. While the facilitated transfer of Li+ ion was not observed across the water/1,2‐dichloroethane interface, the facilitated transfers were observed by formation of 1 : 1 (metal:ionophore) complex for Na+, K+, and Rb+ ions except for Cs+ ion. In the case of Cs+ a 1 : 2 (metal:ionophore) complex was obtained from its special electrochemical response to the variation of ligand concentrations in the organic phase. The logarithms of the complex association constants, for facilitated transfer of Na+, K+, Rb+, and Cs+, were estimated as 6.52, 7.75, 7.91 (log β1°), and 8.36 (log β2°), respectively.  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