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 62 毫秒
1.
Four different fluoride, chloride, bromide, and nitrate cationic polyelectrolytes were prepared. Their electrolytic conductivity in aqueous dilute solution was investigated. The results show a remarkable increase, with dilution being practically constant down to limiting concentration. The limiting equivalent conductivity Λ° follows the order The distances between neighboring charges calculated from experimental data are in good agreement with those calculated geometrically according to Manning's definition. The results are mainly dependent on the assumption that Br? and NO3? interact in a similar manner with the cationic charges on the polyion. This phenomenon also has been observed in cationic micelles.  相似文献   

2.
Chloride ion activity coefficients in aqueous solutions of poly(allylamine) hydrochloride (PAA · HCl) have been determined both in the absence and the presence of simple salts. Without added salt, the activity coefficient depends on the polymer concentration. With added salt, the binding of added counterions by PAAH+ is evaluated from the release of chloride ion. The extent of interaction between counterions and PAAH+ at a given polymer concentration decreases in the order SO ? ClO > NO > Cl? > Br? > I?. This order of counterion selectivity agrees with the previous estimation of potentiometric titrations. The result shows that the hydration of the counterion, as well as its charge, plays an important part in counterion binding to the polyion.  相似文献   

3.
The Unimolecular mass spectrometric fragmentations of the molecular ions of 1,3-diphenylpropane, 1-(7-cycloheptatrienyl)-2-phenylethane and the 1-phenyl-2-tolylethanes and their [d5]phenyl analogues have been investigated by metastable ion techniques and measurements of ionization and appearance energies. By comparing the formation of [C7H7]+, [C7H8]+?, [C8H8]+? and [C8H9]+ it is shown that the molecular ions of the four diaryl isomers do not undergo ring expansion reactions of the aromatic nuclei prior to these fragmentations. Conversely, the molecular ions of the cycloheptatrienyl isomer suffer in part a contraction of the 7-membered ring. From these results and from the measured ionization and appearance energies lower limits to the activation energies of these skeletal isomerizations have been estimated yielding E > 33±5 kcal mol?1 formonoalkylbenzene, E > 20 2±5 kc mol?1 for 7-alkylcycloheptatriene and E > 40±5 kcal mol?1 for dialkylvbenzene positive radical ions. Upper limits can be deduced from literature evidence yielding E < 45 kcal mol?1 for monoalkylbenzene and E < 53 kcal 4mol?1 for dialkylbenzene positive radical ions. The activation energy thus estimated for monoalkylbenzene is in excellent agreement with the recently calculated value(s) for the toluene ion.  相似文献   

4.
The kinetics of the bromate ion-iodide ion-L-ascorbic acid clock reaction was investigated as a function of temperature and pressure using stopped-flow techniques. Kinetic results were obtained for the uncatalyzed as well as for the Mo(VI) and V(V) catalyzed reactions. While molybdenum catalyzes the BrO-I? reaction, vanadium catalyzes the direct oxidation of ascorbic acid by bromate ion. The corresponding rate laws and kinetic parameters are as follows. Uncatalyzed reaction: r2 = k2[BrO] [I?][H+]2, k2 = 38.6 ± 2.0 dm9 mol?3 s?1, ΔH? = 41.3 ± 4.2 kJmol?1, ΔS? = ?75.9 ± 11.4 Jmol?1 K?1, ΔV? = ?14.2 ± 2.9 cm3 mol?1. Molybdenum-catalyzed reaction: r2 = k2[BrO] [I?] [H+]2 + kMo[BrO] [I?] [ H+]2[M0(VI)], kMo = (2.9 ± 0.3)106 dm12 mol?4 s?1, ΔH? = 27.2 ± 2.5 kJmol?1, ΔS? = ?30.1 ± 4.5 Jmol?1K?1, ΔV? = 14.2 ± 2.1 cm3 mol?1. Vanadium-catalyzed reaction: r1 = kV[BrO] [V(V)], kV = 9.1 ± 0.6 dm3 mol?1 s?1, ΔH? = 61.4 ± 5.4 kJmol?1, ΔS? = ?20.7 ± 3.1 Jmol?1K?1, ΔV? = 5.2 ± 1.5 cm3 mol?1. On the basis of the results, mechanistic details of the BrO-I? reaction and the catalytic oxidation of ascorbic acid by BrO are elaborated. © 1995 John Wiley & Sons, Inc.  相似文献   

5.
The equilibrium constants of the reactions MBr2(s) + Al2Br6(sln) ? MAl2Br8(sln) M = Cr, Mn, Co, Ni, Zn, Cd have been measured at 298 K in toluene. Ni: 0.017 ± 0.0024, Co: 0.54 ± 0.07, Zn: 1.5 ± 0.2, Mn: 2.1 ± 0, 7, Cr: 2.2 ± 1, Cd: 7 ± 5. They are compared with literature values of the equilibrium constants of analogous reactions in the gas phase MX2(s) + Al2X6(g) ? MAl2X8(g), X = Cl, Br. For CoAl2Br8(sln) the temperature dependence of the equilibrium constant yielded ΔfH = ?9.4 ± 1 kJ mol?1 and ΔfS = ?39.5 ± 3 J mol?1 K?1 while literature values for CoAl2Br8(g) are ΔfH = 42.4 ± 2 kJ mol?1 and ΔfS = 42.9 ± 2 J mol?1 K?1. The solubility of Al2Br6 in toluene as well as its enthalpy of dissolution have been measured in order to evaluate ΔH° and ΔS° of the solvation of Al2Br6(g) and CoAl2Br8(g) in toluene by a thermodynamic cycle. Solvation of Al2Br6(g): ΔH = ?72.7 ± 1 kJ mol?1, ΔS = ?139.6 ± 4 J mol?1 K?1, solvation of CoAl2Br8(g): ΔH = ?124.5 ± 4kJ mol?1, ΔS = ?222 ± 9J mol?1 K?1. Thus, CoAl2Br8 interacts more strongly with the solvent toluene than Al2Br6 does.  相似文献   

6.
A new two‐step route toward the synthesis of polymeric ionic liquid microgel particles is presented. In the first step, hydrophilic microparticles were prepared by the concentrated emulsion polymerization of the ionic liquid 1‐vinyl‐3‐ethylimidazolium bromide in the presence of small amounts of N,N‐dimethylenebisacrylamide as a crosslinking agent. In the second step, the bromide anion was exchanged in water with different anions such as BF, CF3SO, (CF3SO2)2N?, (CF3CF2SO2)2N?, and dodecylbenzenesulfonate, and this resulted in the coagulation of the microparticles, which were easily recovered by filtration. The obtained polymeric ionic liquid microparticles could be swollen in a very broad range of organic solvents, including apolar organic solvents. As an application, glucose oxidase was encapsulated inside polymeric ionic liquid microparticles, which were used in an amperometric biosensor. The response of the biosensor showed excellent values that strongly depended on the nature of the polymeric ionic liquid counteranion in the order of Br? > BF > (CF3SO2)2N?. © 2006 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 44: 3958–3965, 2006  相似文献   

7.
Kinetics of the complex formation of chromium(III) with alanine in aqueous medium has been studied at 45, 50, and 55°C, pH 3.3–4.4, and μ = 1 M (KNO3). Under pseudo first-order conditions the observed rate constant (kobs) was found to follow the rate equation: Values of the rate parameters (kan, k, KIP, and K) were calculated. Activation parameters for anation rate constants, ΔH(kan) = 25 ± 1 kJ mol?1, ΔH(k) = 91 ± 3 kJ mol?1, and ΔS(kan) = ?244 ± 3 JK?1 mol?1, ΔS(k) = ?30 ± 10 JK?1 mol?1 are indicative of an (Ia) mechanism for kan and (Id) mechanism for k routes (‥substrate Cr(H2O) is involved in the k route whereas Cr(H2O)5OH2+ is involved in k′ route). Thermodynamic parameters for ion-pair formation constants are found to be ΔH°(KIP) = 12 ± 1 kJ mol?1, ΔH°(K) = ?13 ± 3 kJ mol?1 and ΔS°(KIP) = 47 ± 2 JK?1 mol?1, and ΔS°(K) = 20 ± 9 JK?1 mol?1.  相似文献   

8.
The kinetics of the acqueous-phase reactions of the free radicals ·OH, ·Cl, and SO· with the halogenated acetates, CH2FCOO?, CHF2COO?, CF3COO?, and with CH2ClCOO?, CHCl2COO?, CCl3COO? were investigated. Generally, the reactivity decreases with increasing halogen substitution and is in the order k(·OH) > k(SO·) > k(·Cl), but there is no general relation between the effect on reactivity of chlorine and fluorine substitution. © 1995 John Wiley & Sons, Inc.  相似文献   

9.
Gas‐phase reactions of ozone with two butenes (1‐butene and isobutene) and two methyl‐substituted butenes (2‐methyl‐1‐butene and 3‐methyl‐1‐butene) have been studied in an indoor chamber at 295–351 K. The O3 concentrations were monitored by Model 49C‐Ozone analyzer. The butene concentrations were measured by gas chromatography–flame ionization detector. The Arrhenius expressions of k=3.50×10?15e(?1756±84)/T cm3 molecule?1 s?1, k=3.39×10?15e(?1697±52)/T cm3 molecule?1 s?1, k=6.18×10?15e?(1822±80)/T cm3 molecule?1 s?1, and k=7.24×10?14e?(2741±139)/T cm3 molecule?1 s?1 were obtained for the ozonolysis reactions of 1‐butene, isobutene, 2‐methyl‐1‐butene, and 3‐methyl‐1‐butene, respectively. Both the reaction rate constant and activation energy obtained in this work are in good agreement with those reported by using different techniques in the literature. © 2011 Wiley Peiodicals, Inc. Int J Chem Kinet 43: 238–246, 2011  相似文献   

10.
The gas phase iodination of cyclobutane was studied spectrophotometrically in a static system over the temperature range 589° to 662°K. The early stage of the reaction was found to correspond to the general mechanism where the Arrenius parameters describing k1 are given by log k1/M?1 sec?1 = 11.66 ± 0.11 – 26.83 ± .31/θ, θ = 2.303RT in kcal/mole. The measured value of E1, together with the fact that E?1 = 1 ± 1 kcal/mole, provides ΔH(c-C4H7.) = 51.14 ± 1.0 kcal/mole, and the corresponding bond dissociation energy, D(c-C4H7? H) = 96.8 ± 1.0 kcal/mole. A bond dissociation energy of 1.8 kcal/mole higher than that for a normal secondary C? H bond corresponds to one half of the extra strain energy in cyclobutene compared to cyclobutane and is in excellent agreement with the recent value of Whittle, determined in a completely different system. Estimates of ΔH and entropy of cyclobutyl iodide are in very good agreement with the equilibrium constant K12 deduced from the kinetic data. Also in good agreement with estimates of Arrhenius parameters is the rate of HI elimination from cyclobutyl iodide.  相似文献   

11.
The ligands (L) bis (2-pyridyl) methane (BPM) and 6-methyl-bis (2-pyridyl)methane (MBPM) form the three complexes CuL2+, CuL, and Cu2L2H with Cu2+. Stability constants are log K1 = 6.23 ± 0.06, log K2 = 4.83 ± 0.01, and log K (Cu2L2H + 2H2+ ? 2 CuL2+) = ?10.99 ± 0.03 for BPM and 4.56 ± 0.02, 2.64 ± 0.02, and ?11.17 ± 0.03 for MBPM, respectively. In the presence of catalytic amounts of Cu2+, the ligands are oxygenated to the corresponding ketones at room temperature and neutral pH. With BPM and 2,4,6-trimethylpyridine (TMP) as the substrate and the buffer base, respectively, the kinetics of the oxygenation can be described by the rate law with k1 = (5.9 ± 0.2) · 10?13 mol l?1 s?1, k2 = (4.0 ± 0.6) · 10?4 mol?1 ls?1, k3 = (1.1 ± 0.1) · 10?12 mol l?1 s?1, and k4 = (9 ± 2) · 10?14 mol l?1 s?1.  相似文献   

12.
Introduction A series of lanthanide sulfide complexes have beenlargely used for ceramics and thin film materials1 andthese complexes could be prepared from the precursorswhich are the compounds containing lanthanide-sulfurbonds.2-4 For instance, the compounds synthesized with[(alkyl)2dtc]-, phen?H2O and lanthanide salts were usedas the volatile precursors for preparing lanthanide sul-fide, its friction properties in lubricant was investigatedin literature 5 and the preparation and propertie…  相似文献   

13.
The interaction of the palladium(II) complex [Pd(hzpy)(H2O)2]2+, where hzpy is 2‐hydrazinopyridine, with purine nucleoside adenosine 5′‐monophosphate (5′‐AMP) was studied kinetically under pseudo‐first‐order conditions, using stopped‐flow techniques. The reaction was found to take place in two consecutive reaction steps, which are both dependent on the actual 5′‐AMP concentration. The activation parameters for the two reaction steps, i.e. ΔH = 32 ±2 kJ mol?1, ΔS = ?168 ±7 J K?1 mol?1, and ΔH = 28 ± 1 kJ mol?1, ΔS = ?126 ± 5 J K?1 mol?1, respectively, were evaluated and suggested an associative mode of activation for both substitution processes. The stability constants and the associated speciation diagram of the complexes were also determined potentiometrically. The isolated solid complex was characterized by C, H, and N elemental analyses, IR, magnetic, and molar conductance measurements. © 2009 Wiley Periodicals, Inc. Int J Chem Kinet 42: 132–142, 2010  相似文献   

14.
Water exchange on hexaaquavanadium (III) has been studied as a function of temperature (255 to 413 K) and pressure (up to 250 MPa, at several temperatures) by 17O-NMR spectroscopy at 8.13 and 27.11 MHz. The samples contained V3+ (0.30–1.53 m), H+ (0.19–2.25 m) and 17O-enriched (10–20%) H2O. The trifluoromethanesulfonate was used as counter-ion, and, contrary to the previously used chloride or bromide, CF3SO is shown to be non-coordinating. The following exchange parameters were obtained: k = (5.0 ± 0.3) · 102 s?1, ΔH* = (49.4 ± 0.8) kJ mol?1, ΔS* = ?(28 ± 2) JK?1 mol?1, ΔV* = ?(8.9 ± 0.4) cm3 mol?1 and Δβ* = ?(1.1 ± 0.3) · 10?2 cm3 mol?1 MPa?1. They are in accord with an associative interchange mechanism, Ia. These results for H2O exchange are discussed together with the available data for complex formation reactions on hexaaquavanadium(III). A semi-quantitative analysis of the bound H2O linewidth led to an estimation of the proportions of the different contributions to the relaxation mechanism in the coordinated site: the dipole-dipole interaction hardly contributes to the relaxation (less than 7%); the quadrupolar relaxation, and the scalar coupling mechanism are nearly equally efficient at low temperature (~ 273 K), but the latter becomes more important at higher temperature (75–85% contribution at 360 K).  相似文献   

15.
The reactions of stabilized carbonium ions of setoglaucin, methyl violet, and ethyl violet with cyanide ions are largely catalyzed by the cationic micelles of cetyltrimethylammonium bromide (CTAB) in aqueous media. Added counterions (anions in this case) have strong inhibitory effects on the CTAB-catalyzed reactions in the following order: N > NO > Br? > Cl? > F? > no salt. The inhibitory effects of the counterions have been attributed to the exchange between added anions and reagent (CN?) in the micellar media. The data have been analyzed by the model schemes, and mathematical formulations were developed. Various parameters associated with the exchange process, such as equilibrium exchange constant, number of surfactant molecules per substrate molecule, number of added anions, and a factor related to the binding of additives to the catalytic micellar aggregates, have been evaluated.  相似文献   

16.
Incorporation of the lipophilic Co(III)-cobyrinate octadecyl-cobester 1 and of its ionic aqua-cyano perchlorate derivative 2 into poly(vinyl chloride)/bis(1-butylpentyl) adipate liquid membranes induces a selectivity, measured potentiometrically, of about 103 for SCN? an NO with respect to CI?, but only of about 4 for ClO vs. CI?. This is in contrast to classical anion-exchanger membranes, which exhibit a selectivity sequence ClO > SCN? ? NO > Cl? in accordance with the Hofmeister, series. The Co(III)-corrins 1 and 2, when components in solvent polymeric membranes, undergo exchange of axial ligands an behave as highly selective carriers fof SCN? and NO.  相似文献   

17.
The kinetics of the reaction of CH3O with NO and the branching ratio for HCHO product formation, obtained as ΓHCHO = (Rate of HCHO formation) / (Rate of CH3O decay), have been studied using a discharge flow reactor. Laser induced fluorescence has been used to monitor the decay of the CH3O radical and the build-up of the HCHO product. Overall rate constants and product branching ratios were measured at room temperature over the pressure range of 0.72–8.5 torr He. Three reaction mechanisms were considered which differed in the routes of HCHO formation: (i) direct disproportionation; (ii) via an energized collision complex; or (iii) both reaction routes. It has been shown that data on the pressure dependence of the overall rate constant are not sufficient to distinguish between these mechanisms. In addition, an accurate value of Γ is required. Analysis of the available experimental data provided 0.0 and about 0.1 as the lower and upper limit for Γ, respectively. Since the rate constants derived for CH3ONO formation were not sensitive to the value assumed for Γ, k = (1.69 ± 0.69) × 10?29 cm6 molecule?2 s?1 and k = (2.45 ± 0.31) × 10?11 cm3 molecule?1 s?1 could be derived. The rate constant obtained for formaldehyde formation when extrapolated to zero pressure is k = (3.15 ± 0.92) × 10?12 cm3 molecule?1 s?1. © 1994 John Wiley & Sons, Inc.  相似文献   

18.
A study was carried out in aqueous solutions using luminescence technique to investigate the effects of pH, salt concentration, and temperature on the polyacrylic acid/uranyl ion (PAA/UO) complex formation as well as competitive phenomena of enhancement and quenching effects on photoexcited state of uranyl ions. It was found that excess of H+ and OH? is not favorable for complexation between uranyl ions and polymer. Added nitrate salts of Na+ and K+ had significant enhancement effect on emission spectra of PAA/UO complex. These results indicated that the metal ion/polymer chain complex collapsed by addition of salts and then complex became more compact with consequent phase separation. No significant effect of temperature on the PAA/UO complex stability has been observed between 25–50 °C. The quenching rate constants obtained from Stern–Volmer plots were found to be in the order of kq(H+) >> kq(K+) > kq(Na+). © 2005 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 43: 2737–2744, 2005  相似文献   

19.
The title cation ( = Ni2L) is formed in a variety of reactions (Schemes 1 and 2) in systems containing Ni2+ and (2-thiolatoethyl)-diphenylphosphine (= L?) in the absence of coordinating anions at Ni2+/L? ratios > 0.5 in apolar or moderately polar media. Solid [Ni2L3]CIO4 and [Ni2L3]BPh4 have been isolated. Job's plots confirm the Ni2L- stoichiometry in solution. 31P-NMR data are consistent with ≥ 97% Ni2L (vs. ? 3% of hypothetical Ni3L) at equilibrium and support the suggested configuration (Fig. 2). The equilibrium between NiL2 + NiL2Br2 and Ni2L + Br? varies with the solvent composition in CH23Cl2/EtOH mixtures. The rate of formation of Ni2L2Br2 from Ni2L and bromide (in high excess) in CH2Cl2 is first-order in [Ni2L]tot but depends on the ratio [Bu4NBr]tot/[Ni2L3 · ClO4]tot, even at a high excess of bromide. This is interpreted by efficient competition in ion-aggregate formation between the small perchlorate concentration introduced as the counterion of Ni2L, and the large excess of bromide.  相似文献   

20.
Sulfoximide and Sulfoximidium Salts – Structures and Hydrogen Bonding In the solid state dimethylsulfoximide ( 1 ) (orthorhombic; space group Pbca; a = 577.8, b = 931.2 and c = 1645.6 pm) makes intermolecular N? H ? N hydrogen bonds. The hydrogen halide salts (CH3)2S(O)NH2+Hal? (( 2 ), Hal??Cl?; ( 4 ), Hal??Br?) reacts with metal halides to yield (CH3)2S(O)NH2+MHal with the complex anions (( 5 ), MHal?SbCl4?; ( 6 ), MHal?SbCl52?; ( 7 ), MHal?SbCl6?; ( 8 ), MHal?SbBr52?; ( 9 ), MHal?AlCl4?). 2 crystallizes from ethanol (96%) as [(CH3)2S(O)NH2+Cl?]2 · H2O ( 3 ). The structures of 3 (monoclinic; space group P21/c; a = 917.0, b = 1344.7, c = 1080.8 pm and β = 103.8°; Z = 10), 4 (orthorhombic; space group Pbcn; a = 1028.9, b = 1132.6, c = 1074.1 pm; Z = 8) and 6 (monoclinic; space group C2/c; a = 2041.1, b = 1101.4, c = 3365.6 pm and β = 153.8°; Z = 8) are determined by X-ray analysis. In 6 Sb is coordinated in a distorted octahedra by 6 Cl in three short (mean 245,5 pm; SbCl3) and three long distances (291 to 299 pm; Cl?). Two of the chloride ions connect the Sb atoms to infinite Sb …? Cl …? Sb chains. Except for 7 and 9 there are bridges between the NH2 groups and the halide ions. The NH valence vibrations are discussed in view of hydrogen bonding.  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