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 93 毫秒
1.
The distribution of nitric acid between an aqueous phase and a solution of THPO, TcHPO and TOPO in benzene was measured at 25°. The apparent stability constants K1' were found to be 17.2±1.0 (M)-2 for THPO.HNO3, 7.0±0.8 (M)-2 for TcHPO.HNO3 and 15.2 ± I.I (M)-2 for TOPO.HNO3. The stoichionetry of the last complex was confirmed in n-octane solution at constant ionic strength in the aqueous phase.  相似文献   

2.
The distribution of TBPO, water and nitric acid has been measured between an aqueous phase and various inert diluents at 25°. This study allowed the determination of the apparent stability constants for two molecular complexes, TBPO . HNO3 (K1=17.5±1.3 (M/l)-2 (benzene); 20.1±0.8 (M/l)-2 (toluene); II±2(M/l)-2 (n-hexane) and TBPO . 2HNO3.H2O (K2=(2.9±0.3) 10-3(M/l)-2 (benzene)).  相似文献   

3.
P.G. David 《Polyhedron》1985,4(3):437-440
Complex formation between copper(II) and bromide in anhydrous methanol was investigated spectrophotometrically. At a constant copper(II) concentration of 3.0 x 10?4M, Cu2+ and CuBr+ are at equilibrium for [Br?] < 1.0 x 10?3M while CuBr+ and CuBr2 exist at equilibrium in the range of [Br?] 2.0 x 10?3 ?40 x 10?3M. An isosbestic point at 235 nm indicated the equilibrium of Cu2+ and CuBr+ while a second isosbestic point at 290 nm showed the equilibrium of CuBr+ and CuBr2. Stability constants for the formation of CuBr+ and CuBr2 (K1, and K2, respectively) were determined as a function of ionic strength in the range 0.01–0.10. Log K1 and log K2 values at zero ionic strength were obtained by extrapolation of the plot of log K vs ionic strength, the values being 3.97 ± 0.01 and 2.31 ± 0.01.  相似文献   

4.
The following reactions: NpO2++4 HCl ? Np4+ + 2 H2O + 12 Cl2 + 3Cl- NpO2++12 Cl2 ? NpO22+ +Cl- 2HCl+O2- ? H2O +Cl- have been examined quantitatively. The reactions were studied in fused LiCl-KCl sparged with gas mixtures of definite compositions. The concentrations of the diffrent neptunium species were measured by absorption spectrophotometry. The values of respective equilibrium constants are: K = (9.3±0.4)· 10-6 atm(-12); K1 = (2.3±0.1)·10-2 atm(-12); k = 103.8 1 mol-1 atm-1 The standard potential of the system NpO22+ / NpO2+ was determinedto be E0 = 0.220 V (vs.standard chlorine electrode).  相似文献   

5.
J. Toullec  J.E. Dubois 《Tetrahedron》1973,29(18):2851-2858
The kinetics of the iodination of acetone, diethylketone and di-isopropylketone in aqueous media ([H2SO4] = 0·1 to 1·0 N; [I2]ao = 10?7 to 10?5M) have been studied by couloamperometry under irreversible conditions. At these concentrations the rates of formation of the enol and of its iodination are similar. The general equation, which assumes the steady state approximation for the enol, is applicable, and is used to separate the rate constants of enolisation (k1) and the apparent enol iodination rate constant (kIII2 = KEk2I2). For acetone, the value given by Schwarzenbach for the enol equilibrium constant (KE = 2·5 x 10?6) leads to an elementary rate constant for the addition of iodine to the enol (k2I2 = 6·5 x 106 M?1s?1). This value is not, however, consistent with kI2 = 1·5 x 108 M?1s?1, the rate constant for the iodination of the corresponding ether 2-ethoxypropene.  相似文献   

6.
The primary redox reactions for solid-state ion-selective electrodes prepared from electronically semiconducting salts of 7,7,8,8-tetracyanoquinodimethane (tcnq) can be identified by considering the redox properties of their constituent ions or molecules. Three different processes involving the couples, Mn+/M0, 2tcnqo/(tcnq-)2 and (tcnq-)2/2tcnq2- are possible depending on salt composition. Ionic product values determined by potentiometric and atomic absorption methods are in excellent agreement for several such salts; Ks(K2tcnq2)=5.8±1.2·10-11(pot.), 1.7±1·10-11 (a.a.s.); Ks(Cdtcnq2) = 3.0±0.5·10-9 (pot.), 2.9±0.3·10-9(a.a.s.); Ks(Pbtcnq2) = 1.3±0.3·10-10 (pot.), 0.96±0.2·10-10(a.a.s.); and indicate that the lower activity limit for electrode response is controlled by the solubility of the sensor material itself. Comparisons of predicted and observed standard electrode potentials provide quantitative support for an ion-exchange mechanism of interference. The behaviour of electrodes prepared from Cu2tcnq2 (copper(I)) and Cutcnq2 (copper(II)) is explained on the basis of an interference mechanism and considerations of solid-state equilibria.  相似文献   

7.
J.E. Dubois  J. Toullec 《Tetrahedron》1973,29(18):2859-2866
The kinetics of the bromination and chlorination of acetone, diethylketone and di-isopropylketone (bromination only) have been studied at [X2]ao ≈ 10?7 to 10?5 M; the apparent rate constants kIIX2 = KEk2X2 (where KE is the keto-enol equilibrium constant) for iodination, bromination and chlorination are approximately equal. This result is attributed to diffusion-controlled kinetics. The order of magnitude of such a limiting rate constant, 109 M?1s?1 calculated from Smoluchowski's equation, leads to new values for KE in solution (1·5 x 10?8 for acetone) much smaller than those in the literature. The rate constants derived for enol ketonisation are then in good agreement with those from proton addition to the corresponding enol ethers.  相似文献   

8.
The macrocycle-mediated flux of Hg2+ individually and of Hg2+ and Mn+ (Mn+ = K+, Tl+, Ag+, Sr2+, Cd2+, or Pb2+) in cation mixtures has been measured at 25°C in a I M HNO3CHCl31 M HNO3 liquid membrane system. Of the ten macrocycles used, 18-crown-6(18C6), dicyclohexano-18-crown-6(DC18C6), and 21-crown-7(21C7)were most effective in transporting Hg2+ individually. Relative cation fluxes in the metal ion mixtures correlated well with relative log K values for cation--macrocycle interaction and with relative partition coefficients for the distribution of the resulting complex between the aqueous and organic phases  相似文献   

9.
The kinetics of the aquation of (H2O)5Cr(O2CCCl3)2+ have been examined at 35–55°C and 1.00M ionic strength with [H+] = 0.01?1.00M. The reaction follows the rate equation -d ln [Crtotal]/dt = (a[H+]?1 + b + c[H+])/(1 + d[H+]), where [Crtotal] is the stoichiometric concentration of the complex. At 45°C a = (1.41 ± 0.03) × 10?7M/s, b = (1.66 ± 0.02) × 10?5 s?1, c = (7.0 ± 0.8) × 10?5M?1·S?1 and d = 2.3 ± 0.3M?1. Two mechanisms consistent with this rate law are discussed, with evidence being presented in favor of an ester hydrolysis mechanism involving steady-state intermediates. Equilibrium and activation parameters were determined.  相似文献   

10.
A laser photolysis–long path laser absorption (LP‐LPLA) experiment has been used to determine the rate constants for H‐atom abstraction reactions of the dichloride radical anion (Cl2) in aqueous solution. From direct measurements of the decay of Cl2 in the presence of different reactants at pH = 4 and I = 0.1 M the following rate constants at T = 298 K were derived: methanol, (5.1 ± 0.3)·104 M−1 s−1; ethanol, (1.2 ± 0.2)·105 M−1 s−1; 1‐propanol, (1.01 ± 0.07)·105 M−1 s−1; 2‐propanol, (1.9 ± 0.3)·105 M−1 s−1; tert.‐butanol, (2.6 ± 0.5)·104 M−1 s−1; formaldehyde, (3.6 ± 0.5)·104 M−1 s−1; diethylether, (4.0 ± 0.2)·105 M−1 s−1; methyl‐tert.‐butylether, (7 ± 1)·104 M−1 s−1; tetrahydrofuran, (4.8 ± 0.6)·105 M−1 s−1; acetone, (1.41 ± 0.09)·103 M−1 s−1. For the reactions of Cl2 with formic acid and acetic acid rate constants of (8.0 ± 1.4)·104 M−1 s−1 (pH = 0, I = 1.1 M and T = 298 K) and (1.5 ± 0.8) · 103 M−1 s−1 (pH = 0.42, I = 0.48 M and T = 298 K), respectively, were derived. A correlation between the rate constants at T = 298 K for all oxygenated hydrocarbons and the bond dissociation energy (BDE) of the weakest C‐H‐bond of log k2nd = (32.9 ± 8.9) − (0.073 ± 0.022)·BDE/kJ mol−1 is derived. From temperature‐dependent measurements the following Arrhenius expressions were derived: k (Cl2 + HCOOH) = (2.00 ± 0.05)·1010·exp(−(4500 ± 200) K/T) M−1 s−1, Ea = (37 ± 2) kJ mol−1 k (Cl2 + CH3COOH) = (2.7 ± 0.5)·1010·exp(−(4900 ± 1300) K/T) M−1 s−1, Ea = (41 ± 11) kJ mol−1 k (Cl2 + CH3OH) = (5.1 ± 0.9)·1012·exp(−(5500 ± 1500) K/T) M−1 s−1, Ea = (46 ± 13) kJ mol−1 k (Cl2 + CH2(OH)2) = (7.9 ± 0.7)·1010·exp(−(4400 ± 700) K/T) M−1 s−1, Ea = (36 ± 5) kJ mol−1 Finally, in measurements at different ionic strengths (I) a decrease of the rate constant with increasing I has been observed in the reactions of Cl2 with methanol and hydrated formaldehyde. © 1999 John Wiley & Sons, Inc. Int J Chem Kinet 31: 169–181, 1999  相似文献   

11.
《Polyhedron》1988,7(7):573-574
The following equations have been developed to estimate trivalent actinide and lanthanide carbonate stability constants: log B10 = −6.128+35.206R−21.557R2; log B20 = 14.797+7.945R−10.304R2, where B10 = aMCO3+/(aM3+aCO32−, B20 = aM(CO3)2(aM3+(aCO32−), aX is the activity of ion X, M indicates a metal ion and R is the effective ionic radius of the metal ion in six-fold coordination. These equations describe carbonate stability constants for cerium, europium and ytterbium at zero ionic strength. Constants at zero ionic strength were estimated from experimental determinations made in 0.68 molal NaClO4 by accounting for medium effects.  相似文献   

12.
The kinetics and mechanism by which monochloramine is reduced by hydroxylamine in aqueous solution over the pH range of 5–8 are reported. The reaction proceeds via two different mechanisms depending upon whether the hydroxylamine is protonated or unprotonated. When the hydroxylamine is protonated, the reaction stoichiometry is 1:1. The reaction stoichiometry becomes 3:1 (hydroxylamine:monochloramine) when the hydroxylamine is unprotonated. The principle products under both conditions are Cl, NH+4, and N2O. The rate law is given by ?[d[NH2Cl]/dt] = k+[NH3OH+][NH2Cl] + k0[NH2OH][NH2Cl]. At an ionic strength of 1.2 M, at 25°C, and under pseudo‐first‐order conditions, k+= (1.03 ± 0.06) ×103 L · mol?1 · s?1 and k0=91 ± 15 L · mol?1 · s?1. Isotopic studies demonstrate that both nitrogen atoms in the N2O come from the NH2OH/NH3OH+. Activation parameters for the reaction determined at pH 5.1 and 8.0 at an ionic strength of 1.2 M were found to be ΔH? = 36 ± 3 kJ · mol–1 and Δ S? = ?66 ± 9 J · K?1 · mol?1, and Δ H? = 12 ± 2 kJ · mol?1 and Δ S? = ?168 ± 6 J · K?1 · mol?1, respectively, and confirm that the transition states are significantly different for the two reaction pathways. © 2005 Wiley Periodicals, Inc. Int J Chem Kinet 38: 124–135, 2006  相似文献   

13.
The protonation constants and solubilities of three complexons [ethylenediamine-N,N′-disuccinic acid (EDDS), ethylene glycol bis(2-aminoethyl ether)-N,N,N′,N′-tetraacetic acid (EGTA) and 1,2-cyclohexanediamine-N,N,N′,N′-tetraacetic acid (CDTA)] are reported in aqueous solutions of NaCl with different ionic strength values (0 ≤ I ≤ 4.8 mol·L?1) and, in the case of CDTA, in (CH3)4NCl (0.1 ≤ I ≤ 2.7 mol·L?1). The dependence on ionic strength of the protonation constants of these three complexons and four other complexons that were previously reported (NTA, EDTA, DTPA and TTHA), is analyzed in NaCl solution; the ionic strength influences quite strongly the protonation constants (as an example for CDTA, log10 K 1 = 10.54 and 9.25 at I = 0.1 and 1 mol·L?1, respectively), while the effect of (CH3)4NCl concentration is lower. Based on the total solubility S T and the protonation constant data at different salt concentrations, the solubility of the neutral species S 0 and the solubility products K S0 are obtained. The Setschenow coefficients k m and the solubility values S 0 0 in pure water are also reported (S 0 0  = 0.55, 0.21 and 0.75 mmol·kg?1 for EDDS, EGTA and CDTA, respectively). The dependence of the protonation constants on ionic strength is also interpreted in terms of ion pair formation, and the formation constants of Na+ species are reported.  相似文献   

14.
The composition of complexes formed upon the extraction of UVI and ThIV nitrates with O-n-nonyl(N,N-dibutylcarbamoylmethyl) methyl phosphinate (L) from solutions of nitric acid without additional solvent was determined by 31P NMR spectroscopy. The structures of the complexes formed were studied by IR spectroscopy. Uranium(VI) is extracted from 3 and 5 M solutions of HNO3 as the [UO2(L)2(NO3)2] complex, while thorium(IV) is extracted from 5 M HNO3 as the [Th(L)3(NO3)3]+·NO 3 complex. In both cases, ligand L has bidentate coordination. Ligand L contacts with 3 and 5 M nitric acid to form adducts L·HNO3 and L· (HNO3)2, respectively. Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 11, pp. 2460–2464, November, 2005.  相似文献   

15.
The Crystal Structures of [Cu2Cl2(AA · H+)2](NO3)2 and [AA · H+]Picr? (AA · H+ = Allylammonium; Picr? = Picrat) By an alternating current electro synthesis the crystal-line π-complex [Cu2Cl2(AA · H+)2](NO3)2 has been obtained from CuCl2 · 2H2O, allylamine (AA), and HNO3 in ethanolic solution. X-ray structure analysis revealed that the compound crystallized in the monoclinic system, space group P21/a, a = 7.229(3), b = 7.824(3), c = 26.098(6) Å, γ = 94.46(5)°, Z = 4, R = 0.025 for 2 023 reflections. The crystal structure is built up of CunCln chains which are connected by π-bonding bidentate AA · H+ …? ON(O)O …? H+ · AA units. For comparision with the above complex the structure of [AA · H+]Picr? (Picr? = picrate anion) is also reported.  相似文献   

16.
The complex formation of Fe3+ with o-methyl benzamide oxime was studied spectrophotometrically in methanol solution. The stepwise process gives complexes 1∶1, 1∶2 and 1∶3. The formation constants are lgK 1 = lg β1 = 1,88 ± 0,12, lgK 2 = 3,53 ± 0,2, lgK 3 = 4,96 ± 0,2, lg β2 = 1,65 ± 0,32 and lg β3 = 1,43 ± 0,4, whereK 3 = β1 · β2 · β3. All measurements were carried out at 25°C and an ionic strength μ=1.  相似文献   

17.
Thermal decomposition of neat TBP, acid-solvates (TBP·1.1HNO3, TBP·2.4HNO3) (prepared by equilibrating neat TBP with 8 and 15.6?M nitric acid) with and without the presence of additives such as uranyl nitrate, sodium nitrate and sodium nitrite, mixtures of neat TBP and nitric acid of different acidities, 1.1?M TBP solutions in diluents such as n-dodecane (n-DD), n-octane and isooctane has been studied using an adiabatic calorimeter. Enthalpy change and the activation energy for the decomposition reaction derived from the calorimetric data wherever possible are reported in this article. Neat TBP was found to be stable up to 255?°C, whereas the acid-solvates TBP·1.1HNO3 and TBP·2.4HNO3 decomposed at 120 and 111?°C, respectively, with a decomposition enthalpy of ?495.8?±?10.9 and ?1115.5?±?8.2?kJ?mol?1 of TBP. Activation energy and pre exponential factor derived from the calorimetric data for the decomposition of these acid-solvates were found be 108.8?±?3.7, 103.5?±?1.4?kJ?mol?1 of TBP and 6.1?×?1010 and 5.6?×?109?S?1, respectively. The thermochemical parameters such as, the onset temperature, enthalpy of decomposition, activation energy and the pre-exponential factor were found to strongly depend on acid-solvate stoichiometry. Heat capacity (C p ), of neat TBP and the acid-solvates (TBP·1.1HNO3 and TBP·2.4HNO3) were measured at constant pressure using heat flux type differential scanning calorimeter (DSC) in the temperature range 32?C67?°C. The values obtained at 32?°C for neat TBP, acid-solvates TBP·1.1HNO3 and TBP·2.4HNO3 are 1.8, 1.76 and 1.63?J?g?1?K?1, respectively. C p of neat TBP, 1.82?J?g?1?K?1, was also measured at 27?°C using ??hot disk?? method and was found to agree well with the values obtained by DSC method.  相似文献   

18.
The formation of VC-SO2 and VC-(SO2)2 complexes in liquid mixtures of vinyl chloride (VC) and sulphur dioxide has been shown by (a) the freezing point composition diagram and (b) chemical shifts in the PMR spectrum of VC over the complete composition range. It is postulated that SO2 can associate with the CC bond and the Cl atom. These complexes may be involved in the copolymerization and influence the composition and stereochemistry of the product. PMR spectra of VC-SO2-ethane(E) mixtures with [SO2] ? [E] ? [VC] gave Kv = 2·0 ± 0·5, 1·5 ± 0·1 and 1·1 ± 0·3 at 232·6, 272·6 and 301·3 K with ΔHf0Hf = ?6·6 ± 1·4 kJ mol?1 for the VC -(SO2)2 complex. The chemical shift of the trans β-proton was twice that of the other two protons. indicating that SO2 adopts an asymmetric orientation to the double bond.  相似文献   

19.
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   

20.
Formation constants for yttrium and rare earth element (YREE) chloride complexation have been measured at 25°C by examining the influence of medium (NaClO4 and NaCl) on YREE complexation by fluoride ions and methyliminodiacetate (MIDA). YREE chloride complexation constants Clβ1(M) obtained in this work using dissimilar procedures are in good agreement and indicate that, at constant temperature and ionic strength, Clβ1(M) does not vary significantly across the fifteen-member series of elements. The ionic strength μ dependence of YREE chloride formation constants between 0 and 6 molar ionic strength can be written, for all YREE, as ${\text{log}}_{{\text{CI}}} \beta _1 \left( M \right) = \log _{{\text{CI}}} \beta _1^0 \left( M \right) - 3.066\mu ^{0.5} /\left( {1 + 1.727\mu ^{0.5} } \right)$ where Clβ1(M) = [MCl2+][M3+]?1[Cl?1]?1 and logClβ1 o(M) represents the MCl2+ formation constant for all YREE at zero ionic strength: logClβ1 o = 0.65 ± 0.05.  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