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.
To understand the behaviour of Tl/II/ and the bridging group SO 4 2– in radiation induced electron exchange reactions we have investigated the rate constants and mechanisms of reaction of Tl/II/ with Tl/I/ and Tl/III/ in perchloric acid solutions. The results indicated that Tl/II/ is an intermediate in the -ray induced electron exchange process of T1/I/–T1/III/. Sulfate ions at [SO 4 2– ]>-0.02M serve as bridging groups and play an important role in accelerating the T1/II/–T1/I/ reaction. A cooperative effect was found between hydrogen peroxide and sulfate ion at low sulfate ion concentration, [SO 4 2– ]0.02M in perchloric acid solution.  相似文献   

2.
In order to reveal to what extent tritium (3H or T) can be incorporated into hydroxides, the isotope exchange reaction (OT-for-OH exchange reaction) between each alkaline earth metal hydroxide (M(OH)2), where M means alkaline earth metal (M=Ca, Sr or Ba) and HTO water was observed homogeneously at 30 °C under equilibrium after mixing. Consequently, the followings were obtained: (1) a quantitative relation between the electronegativity of each M ion and the ability (of the M ion) incorporating OT into the M hydroxide can be found and the ability is small when the temperature is high, (2) the exchange rate for the OT-for-OH exchange reaction is small when the electronegativity of the M ion in the M hydroxide in great, (3) as for the dissociation of HTO water, it seems that fomula (HTOT++OH) is more predominant than the fomula (HTOH++OT) when the temperature is high and (4) the method used in this work is useful to estimate the reactivity of a certain alkaline material.  相似文献   

3.
Isotope exchange reaction between NaCl-36 and triphenyltin chloride in dioxane-water (8020% w/w) and ethanol-water (9010% w/w) mixed solvents has been studied at 25, 35 and 50 °C. The exchange reaction was found to proceed via a bimolecular SN2, limiting mechanism with reaction rates depending on the solvent used. Inhibition of the exchange in ethanol-water is probably due to solvation of chloride ion through hydrogen bond formation. The rate laws for the exchange reactions are: Re=3.24×109 e–65550/RT [Rh3SnCl] [NaCl] in dioxanewater and Re=6.61×108 e–69600/RT [Ph3SnCl] [NaCl] in ethanol-water, where is the degree of dissociation of NaCl and Re is the rate of exchange in mol l–1 s–1. The activation parameters H*, S* and G* are reported.  相似文献   

4.
Summary The kinetics of OsO4-catalysed oxidation of cyclopentanol, cyclohexanol and cyclooctanol by alkaline hexacyanoferrate(III) have been studied at low [OH] so that the equilibrium between alcohol and alkoxide ion is not unduly shifted towards the latter. The reaction shows a first-order dependence in [OH]. The order of the reaction with respect to cycloalcohol is fractional, indicating the formation of an intermediate complex with OsVIII since the order with respect to hexacyanoferrate(III) ion is zero. The order with respect to OsVIII may be expressed by the equation kobs=a+b[OsVIII]. The analysis of the rate data indicates a significant degree of complex formation between [OsO3(OH)3] and ROH. It was found that the bimolecular rate constant k for the redox reaction between complex and OHk1, the forward rate constant for the formation of alkoxide ion. The activation parameters of these rate constants are reported.  相似文献   

5.
NMR and IR spectroscopy have been used in studying the equilibrium in the reaction of proton transfer from thiocarboxylic acids RCOSH [R=CH3 (a), C6H5 (b) or CH2Cl (c)] to -collidine (d), and also the kinetics of CH/NH proton exchange between protonated -collidine and excess RCOSH. For compoundsa and b, partial protonation of the -collidine is observed; and for compound c, complete protonation. The heat of reaction of proton transfer with the participation of binary acidamine associates is 30 (a) and 45 (b) kJ/mole. The rate of proton exchange decreases and the activation energy E increases with increasing acidity of the RCOSH [E=44 (b) and 88 (c) kJ/mole] and with increasing basicity of the amine (Ed < ETEA), which, in accordance with the orders of reaction that were found for the exchanged components, is due to a mechanism in which the slow stage is proton transfer in the ion pair NH+...SOCR. The thiocarboxylate ion of c is unstable; and after splitting out Cl, it forms the compounds Cl(CH2COS)2 and (CH2COS)2.Translated from Teoreticheskaya i Éksperimental'naya Khimiya, Vol. 21, No. 2, pp. 187–194, March–April, 1985.  相似文献   

6.
The reaction of the azide ion with the carbocation generated in the photolysis of 1,2,2,4,6-pentamethyl-1,2-dihydroquinoline in methanol was studied by pulse (conventional and laser) and steady-state photolysis techniques. The adduct of the azide ion was characterized by 1H NMR spectrum. Experimental results were interpreted taking into account a competition between the addition of methanol and azide ion to the carbocation. The rate constants for the reaction of the azide ion with the carbocation (k Az) were measured at 2—48 °C in a wide range of [N3 ]0 concentrations from 2·10–7 to 0.1 mol L–1 at different ionic strengths () of the solution. The resulting k Az values are more than an order of magnitude lower than those for diffusional-controlled reactions and vary from 3.2·108 ( = 0) to 4.5·106 L mol–1 s–1 ( = 0.8 mol L–1) in the presence of NaClO4 (18 °C). The activation energy of addition of the azide ion to the carbocation is 21 kJ mol–1, which is by 12 kJ mol–1 lower than the activation energy of the reaction of the carbocation with methanol. The features of the reaction under study are discussed from the viewpoint of the structures of carbocations generated in the photolysis of dihydroquinolines.  相似文献   

7.
Cyclic voltammetric reduction of 17 different organic halides including alkyl, allyl, and benzyl chlorides, bromides, and iodides in aprotic media (DMF, DMSO, and acetonitrile, AN), containing 0.1 M Bu4NClO4, have been reported at glassy carbon (gc) and graphite as working electrodes. A single two-electron irreversible and diffusion-limited reduction of the carbon-halogen bond is observed, the reduction potentials ranging from –1.20 to –2.70 V (versus silver wire quasireference electrode) depending on the halide, the solvent, and the electrode. Indirect reduction of these 17 halides, however, is effected at much lower potentials (–0.79 to –0.92 V versus SCE) depending on the experimental conditions by in situ electrogenerated superoxide ion (O 2 ) by CPE. The products have been characterized by TLC, GC, CV, or chemical estimation. The diorganic peroxide and the organic hydroperoxide were the major products. In case of tertiary alkyl halides, however, alkenes predominated, due to basic nature of O 2 in these reactions. These studies indicate sufficient strength of O 2 as a nucleophile or base depending on the experimental conditions.From Elektrokhimiya, Vol. 41, No. 3, 2005, pp. 350–355.Original English Text Copyright © 2005 by Vasudevan.This article was submitted by the author in English.  相似文献   

8.
Reactions of At//+, Ato.H2O, AtCl 2 and AtBr2 with the pseudohalogenides tricyanomethanide and azide are described. Information on the compound formation of astatine with C/CN/ 3 and N 3 could be obtained on the basis of electromigration investigations under variation of the conditions /composition of the electrolyte, pH, exchange reactions of ligands/. For the reaction: [At/H2O/C/CN/3]+C/CN/ 3 [At/C/CN/3/2]+H2O at 301 K and u=0.075 mol.l–1 K2=/675±25/ [1.mol–1] and uo=–/3.50±0.10/×10–4 [cm2.s–1.V–1]. According to this astatine/I/-tricyanomethanide is classified between AtI 2 and At/SCN/ 2 . First investigations in azid-containing systems confirm the formation of astatine/I/-azide-compounds. Their composition is probably At/N3/ 2 . There is no dependence of the ion mobility of astatine/I/-azide in the investigated range on azide concentration which is due to its high stability.  相似文献   

9.
The slow reaction between peroxodisulfate and formate is significantly accelerated by ascorbate at room temperature. The products of this induced oxidation, CO2 and oxalate (C2O2– 4), were analyzed by several methods and the kinetics of this reaction were measured. The overall mechanism involves free radical species. Ascorbate reacts with peroxodisulfate to initiate production of the sulfate radical ion (SO 4), which reacts with formate to produce carbon dioxide radical ion (CO 2) and sulfate. The carbon dioxide radical reacts with peroxodisulfate to form CO2 or self-combines to form oxalate. Competition occurring between these two processes determines the overall fate of the carbon dioxide radical species. As pH decreases, protonation of the carbon dioxide radical ion tends to favor production of CO2.  相似文献   

10.
A kinetic study of the exchange reaction between UO2EDTA complex and Fe(III), at a constant ionic strength of 0.1, over the concentration range of 5×10–3–1×10–2 M of each reactant and pH 4.5–5.5 has been carried out radiometrically. The rate of the exchange process can be expressed by the equation: R=k1[UO2EDTA][Fe]+k2[EDTA][H+]–1. The activation parameters calculated were H*=25.95 kJ mol–1 and S*=0.67 kJ mol–1 K–1.  相似文献   

11.
Calculations were carried out using the semiempirical quantum chemical AMI method for anion radicals (AR) of the perfluoroalkyl halides (RFX): CF3X, CF3CF2X, (CF3)2-CFX, and (CF3)3CX for X=Cl, Br, and I. All the AR's studied are thermally stable. The electron affinity of the perfluoroalkyl halides, and consequently, the thermal stability of their AR's increases in the series from F-methyl to F-tertbutyl halides and from the chlorides to bromides and iodides. During formation of an AR the spin density is preferentially localized on the * orbital of the C–X bond which leads to an increase in the distance between these atoms. Dissociation of the AR of tert-perfluorobutyl iodide to a perfluorocarbanion and an I atom is thermodynamically more favorable than dissociation with formation of a perfluoroalkyl radical and I.Translated from Izvestiya Akademii Nauk SSSR, Seriya Khimicheskaya, No. 5, pp. 1064–1068, May, 1990.  相似文献   

12.
A comparison of proton exchange reactions between OH, SH, and CH acids and the NH groups of trialkylammonium ions showed that regardless of the nature of the acid XH, the mechanism of exchange includes transfer of a proton in the ion pair N-H+ ... X as the slow step. At the fast steps of proton exchange XH- N+H, i.e., molecular exchange with breaking of a hydrogen bond X-H ... N and transfer of a proton along these bonds, differences appear in the properties of XH acids. In the sequence from OH to SH and CH acids, the hydrogen bonds X-H ... N are weakened. As a result of this, in the same sequence the kinetic acidity (k2) decreases but the rate of molecular exchange (kH) increases. The ratio between the values of k2 and kH is inverted when the strong bonds O-H ... N (k2/kH 1) are replaced by weak bonds C-H ... N (k2/kH 1). It was also established that the kinetic stability of the anions increases as the oxygen atoms are replaced by sulfur in the series RCOO < RCOS < R2PSS as a result of the more effective delocalization of the negative charge on the diffuse orbitals of sulfur.Translated from Teoreticheskaya i Éksperimental'naya Khimiya, Vol. 23, No. 4, pp. 471–475, July–August 1987.  相似文献   

13.
Summary The kinetics of the OsVIII-catalysed oxidation of glycols by alkaline hexacyanoferrate(III) ion exhibits zerothorder dependence in [Fe(CN) 6 3– ] and first-order dependence in [OsO4]. The order with respect to glycols is less than unity, whereas the rate dependence on [OH] is a combination of two rate constants; one independent of and the other first-order in [OH]. These observations are commensurate with a mechanism in which two complexes, [OsO4(H2O)G] and [OsO4(OH)G]2–, are formed either from [OsO4(H2O)(OH)] or [OsO4(OH)2]2– and the glycol GH, or by [OsO4(H2O)2] and [OsO4(H2O)(OH)] and the glycolate ion, G, which is in equilibrium with the glycol GH through the reaction between GH and OH. Hence there is an ambiguity about the true path for the formation of the two OsVIII-glycol complexes. A reversal in the reactivity order of glycols in the two rate-determining steps, despite the common attack of OH ion on the two species of OsVIII-complexes, indicates that the two complexes are structurally different because S changes from the negative (corresponding to k11) to positive (related to k2).  相似文献   

14.
Zusammenfassung Ausgehend von der klassischen Theorie, wurde die Bromat-Bromid-Ascorbinsäure-Reaktion untersucht und festgestellt, daß die Grundreaktion zwischen BrO3 und Ascorbinsäure unmeßbar langsam vor sich geht. Durch die Messung der Reaktionszeit des Systems erhielt man für die isolierte Bromat-Bromid-Reaktion eine Geschwindigkeitsgleichung vierter Ordnung. Die Geschwindigkeitskonstante der Reaktion, ungefähr 200 I3 · Mol–3 · min–1, stimmt mit den Angaben der Literatur gut überein. Die Geschwindigkeitskonstante der durch Vanadin katalysierten Reaktion beträgt 7,2 · 107 I4 · Mol–4 · min–1. Zwischen der Vanadinkonzentration und der Reaktionszeit besteht ein linearer Zusammenhang, der kinetisch nachgewiesen werden kann. Dem vorgeschlagenen Reaktionsmechanismus zufolge geht die Elektronenübertragung im Bromat-Bromid-Ascorbinsäure-System in einer Säure-Basen-Reaktions-Reihe im Lewisschen Sinne vor sich, in der als Katalysator nicht nur Bromid und Vanadin, sondern auch H+-Ion wirkt.
Summary Proceeding from the classic kinetic theory, a study was made of the bromate-bromide-ascorbic acid reaction and it was found that the fundamental reaction between BrO3 and ascorbic acid is immeasurably slow. By determining the reaction time of the system, a velocity equation of the fourth order was obtained for the isolated bromate-bromide reaction. The velocity constant of the reaction, approximately 200 liter3 · mol–3 · min–1, agress well with the figures given in the literature. The velocity constant of the reaction catalyzed by vanadium is 7,2·107 liter4 · mol–4 · min–1. There is a linear relation between the vanadium concentration and the reaction time; this can be shown kinetically. According to the proposed reaction mechanism the electron transfer in the bromate-bromide-ascorbic acid system occurs in the Lewis sense in an acid-base reaction series, in which not only bromide and vanadium function as catalyst but also the H ion.
  相似文献   

15.
The conductances of eleven 1:1 salts have been measured at 50°C in N,N-dimethylmethanesulfonamide (DMMSA) for electrolyte concentrations of 1.2–55.0×10–4 mol-dm–3. The conductance data were analyzed using the equation of Lee and Wheaton. Calculations for different values of the distance parameter R indicate that all salts studied are only slightly associated in DMMSA. Association was somewhat greater for the trimethylphenylammonium halides than for the tetraalkylammonium salts. Ionic limiting molar conductances were estimated using the tris (isopentyl) butylammonium tetraphenylborate approximation. The markedly smaller value for o (Na+) compared to the values for o (Br) and o (I) indicates that the sodium ion is probably more extensively solvated than the halide ion. In general, it appears that DMMSA (dielectric constant=80.31 at 50°C) is similar in its solvent properties to dipolar aprotic heterocyclic solvents such as 2-cyanopyridine and 3-methyl-2-oxazolidone which have similar dielectric constants.  相似文献   

16.
Within the framework of the Hartree-Fock-Roothaan Method, using double- basis sets 3-21++G and (6-31-H-G//3-21++G), the minimum energy paths (MEPs) have been calculated for reactions of nucleophilic addition of the hydride ion H to the methylacetylene molecule: CH3-CCH+H[CH3-CH=CH] (1) CH3-CCH+H[CH3-C=CH2] (2). It has been established that the activation energy for reaction (2) is 7.02 kJ/mole lower than for reaction (1). An analysis has been made of the character of electron density distribution along the MEP of each reaction. It has been shown that distortion of geometry of the reactants plays an important role in intensifying the interaction of the frontier orbitals. The reasons for nonfulfillment of Markownikoff's rule for these reactions have been determined. The results from the calculations are compared with calculations reported in the literature for the related reaction of nucleophilic addition of the hydride ion H to the acetylene molecule: HCCH+H[CH2=CH].Translated from Teoreticheskaya i Éxperimental'naya Khimiya, Vol. 21, No. 3, pp. 303–309, May–June, 1985.  相似文献   

17.
Water exchange on Mn centers in proteins has been modeled with density functional theory using the B3LYP functional. The reaction barrier for dissociative water exchange on [MnIV(H2O)2(OH)4] is only 9.6 kcal mol–1, corresponding to a rate of 6×105 s–1. It has also been investigated how modifications of the model complex change the exchange rate. Three cases of water exchange on Mn dimers have been modeled. The reaction barrier for dissociative exchange of a terminal water ligand on [(H2O)2(OH)2MnIV(-O)2MnIV(H2O)2(OH)2] is 8.6 kcal mol–1, while the bridging oxo group exchange with a ring-opening mechanism has a barrier of 19.2 kcal mol–1. These results are intended for interpretations of measurements of water exchange for the oxygen evolving complex of photosystem II. Finally, a tautomerization mechanism for exchange of a terminal oxyl radical has been modeled for the synthetic O2 catalyst [(terpy)(H2O)MnIV(-O)2MnIV(O)(terpy)]3+ (terpy=2,2:6,2-terpyridine). The calculated reaction barrier is 14.7 kcal mol–1.Contribution to the Björn Roos Honorary Issue  相似文献   

18.
Summary The oxidation of formate ion by alkaline osmium tetroxide, such that [HCO inf2 p– ] [OsVII], exhibits first-order dependence in [OsVII], an order less than unity in [HCO inf2 p– ], and zero-order in [OH]. HCO2– reacts as an ion-pair formed with an alkali metal ion and [OsO4(OH)2]2– is the reactive species of OsVII. The formation of an intermediate OsVII-HCO2M complex is substantiated by the rapid-scan spectra of the reaction mixture. Anions (Cl, ClO inf4 p– ) have no effect on the rate. The close agreement between the observed k H/k D = 7.1 and the theoretically calculated value (7.0), based on the stretching frequencies of C-H and C-D bonds in the free molecule, indicates an outer-sphere mechanism.Author to whom all correspondence should be directed.  相似文献   

19.
From conductometric and UV-VIS spectrophotometric studies of the reaction between 18-crown-6 (L) and dichloropicric acid (HA) in dry and water saturated 1,2-dichloroethane, it has been concluded that formation of a 1:1 homoconjugate HA 2 accompanies the simple protonation of L, viz, L+HALH+A and L+2HALH+HA 2 . The electrolytes LH+A and LH+HA 2 are extensively, or practically completely dissociated in both solvents under the experimental conditions. The specie LH+A appears to be a contact ion pair in DCE. The stability constant of HA 2 in the dry solvent, 5.7×103 mol–1-cm3, is some 102.4 times that in propylene carbonate reflecting the difference in H-bond accepting capacity of the two solvents. Hydration of HA, A and HA 2 in wet dichloroethane is negligible or slight. As expected, LH+ is rather strongly hydrated, the ratio of the hydration constants of LH+ and L being about 1×101.  相似文献   

20.
The ion exchange processes of (OAc) and (OAc) proceeding in shell-core inorganic ion exchanger Ti (HPO4)2·1/2H2O has been studied and the diffusion equation whose boundary conditions are satisfied by a shell-core model was solved. Based on the equation solved and experimental data, the diffusion coefficients corresponding to the exchange process (OAc) and Li+–H+ (OAc) at 17°C are found to be 7.7×10–9 and 6.2×10–8 cm2 s–1 and the activation energies 3.4×104 and 5.0×103 J mol–1, respectively. Compared to the gel type of styrene-divinylbenzene strong acid exchanger with 20% cross linking, it can be concluded that the rate of or exchange is 3.5 times faster than that in the organic exchanger.TIP was obtained from the Salt Lake Institute of the Academy of Science of China.  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