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 46 毫秒
1.
J.G. Leipoldt  H. Meyer 《Polyhedron》1985,4(9):1527-1531
The reaction of Cl?, Br?, I?, Co(CN)63? and NCS? with meso-tetrakis (p-trimethylammoniumphenyl)porphinatodiaquorhodate(III), [RhTAPP(H2O)2]5+, has been studied at 15, 25 and 35°C in 0.1 M [H+] with μ = 1.00 M (NaNO3). The value of the acidity constant, Kal, at 25°C is 4.39 × 10?9 M. The reactions are first order in anion concentration up to 0.9 M. The values of the stability constants, K1, and the second order rate constants, k1, for the reaction with Cl?, Br?, I?, Co(CN)63? and NCS? are respectively 0.23 M?1 and 2.5 × 10?3 M?1 s?1, 1.1 M?1 and 6.92 × 10?3 M?1 s?1, 40.0 M?1 and 17.0 × 10?3 M?1 s?1, 550 M?1 and 20.0 × 10?3 M?1 s?1, 3400 M?1 and 20.9 × 10?3 M?1 s?1. The porphine greatly labilizes the Rh(III). There has been about a 500-fold increase in the rate constant for substitution compared to that of [Rh(NH3)5H2O]3+. The substitution rates are however about the same as for [Rh(TPPS)(H2O)2]3?, indicating that the overall charge on the complex plays only a minor role. The kinetic results indicate that dissociative activation is occurring in these reactions.  相似文献   

2.
A jet-stream kinetic technique and the resonance fluorescence method applied to detection of iodine atoms were used to measure the rate constants of the reactions of the IO· radical with the halohydrocarbons CHFCl-CF2Cl (k = (3.2 ± 0.9) × 10?16 cm3 molecule s?1) and CH2ClF (k = (9.4 ± 1.3) × 10?16 cm3 molecule s?1), the hydrogen-containing haloethers CF3-O-CH3 (k = (6.4 ± 0.9) × 10?16 cm3 molecule s?1) and CF3CH2-O-CHF2 (k = (1.2 ± 0.6) × 10?15 cm3 molecule s?1), and hydrogen iodide (k = (1.3 ± 0.9) × 10?12 cm3 molecule s?1) at 323 K.  相似文献   

3.
An analysis of the former works devoted to the reactions of I(III) in acidic nonbuffered solutions gives new thermodynamic and kinetic information. At low iodide concentrations, the rate law of the reaction IO + I? + 2H+ ? IO2H + IOH is k+B [IO][I?][H+]2k?B [IO2H][IOH] with k+B = 4.5 × 103 M?3s?1 and k?B = 240 M?1s?1 at 25°C and zero ionic strength. The rate law of the reaction IO2H + I? + H+ ? 2IOH is k+C [IO2H][I?][H+] – k?C [IOH]2 with k+C = 1.9 × 1010 M?2s?1 and k?C = 25 M?1s?1. These values lead to a Gibbs free energy of IO2H formation of ?95 kJ mol?1. The pKa of iodous acid should be about 6, leading to a Gibbs free energy of IO formation of about ?61 kJ mol?1. Estimations of the four rate constants at 50°C give, respectively, 1.2 × 104 M?3s?1, 590 M?1s?1, 2 × 109 M?2s?1, and 20 M?1 s?1. Mechanisms of these reactions involving the protonation IO2H + H+ ? IO2H and an explanation of the decrease of the last two rate constants when the temperature increases, are proposed. © 2008 Wiley Periodicals, Inc. Int J Chem Kinet 40: 647–652, 2008  相似文献   

4.
The reduction of Fe(CN)5L2? (L = pyridine, isonicotinamide, 4,4′‐bipyridine) complexes by ascorbic acid has been subjected to a detailed kinetic study in the range of pH 1–7.5. The rate law of the reaction is interpreted as a rate determining reaction between Fe(III) complexes and the ascorbic acid in the form of H2A(k0), HA?(k1), and A2? (k2), depending on the pH of the solution, followed by a rapid scavenge of the ascorbic acid radicals by Fe(III) complex. With given Ka1 and Ka2, the rate constants are k0 = 1.8, 7.0, and 4.4 M?1 s?1; k1 = 2.4 × 103, 5.8 × 103, and 5.3 × 103 M?1 s?1; k2 = 6.5 × 108, 8.8 × 108, and 7.9 × 108 M?1 s?1 for L = py, isn, and bpy, respectively, at μ = 0.10 M HClO4/LiClO4, T = 25°C. The kinetic results are compatible with the Marcus prediction. © 2005 Wiley Periodicals, Inc. Int J Chem Kinet 37: 126–133, 2005  相似文献   

5.
Vanadium(II) ions form with the pyridine-2-carboxylate ligand a deep blue, tris-substituted complex absorbing at 660 nm (ε = 7.2 × 103 M?1) cm?1) with a shoulder at 450 nm. Reversible spectroelectrochemistry and cyclic voltammetry were observed for this complex, with E12 = ?0.448 V vs NHE, and ΔSrcθ = ?6 cal · mol?1 · deg?1. Electron transfer kinetics with [CO(en)3]3+ led to k12 = 3100 M?1 s?, ΔH = 12.4 kcal · mol?1 and ΔS = ?0.9 cal · mol?1 · deg?1 (I = 0.10 M). For the related [Co(NH3)6]3+ complex, k13 = 1.9 × 104 M?1 s?1. The self-exchange rate constant and activation parameters were analysed in terms of relative Marcus theory.  相似文献   

6.
The oxidation processes of the radiation-generated, three-electron-bonded intermediates AcMet2 [S??S]+ and AcMet [S??Br] were investigated by pulse radiolysis via their reactions with tryptophan (TrpH). These intermediates were derived from N-acetyl-methionine amide (N-AcMetNH2) and N-acetyl-methionine methyl ester (N-AcMetOMe). The bimolecular rate constant k of the reaction between each intermediate and l-tryptophan (TrpH) was measured. For N-AcMetNH2, k for the reaction of AcMet2 [S??S]+ with TrpH were 3.4?×?108 and 2.2?×?108?dm3?mol?1?s?1 at pH?=?1 and 4.5, respectively. For N-AcMetOMe, k for the reaction of AcMet2 [S??S]+ with TrpH were 4.0?×?108 and 2.8?×?108?dm3?mol?1?s?1 at pH 1 and 4.5, respectively. The rate constants for the intermolecular transformation of Met [S??Br] into TrpH+ or Trp were also estimated. For N-AcMetNH2, k for the reaction of AcMet2 [S??Br] with TrpH were 2.6?×?108 and 3.3?×?108?dm3?mol?1?s?1 at pH 1 and 4.5, respectively. Related mechanisms were discussed.  相似文献   

7.
The variation in the lifetime of flash-excited gaseous benzophenone with pressure and temperature indicates that (1) self-quenching is a relatively inefficient process for the long-lived emission, ksq = 9 × 105 M?1 s?1 (estimated from solution data) at 25°C and 1.2 × 107 M?1 s?1 at 170°C and (2) the lifetime decreases with increasing temperature as a result of photochemical and photophysical decay pathways which have significant activation energies. The importance of diffusion to the walls on lifetime measurements is discussed.  相似文献   

8.
The kinetics of oxidation of tartaric acid (TAR) by peroxomonosulfate (PMS) in the presence of Cu(II) and Ni(II) ions was studied in the pH range 4.05–5.20 and also in alkaline medium (pH ~12.7). The rate was calculated by measuring the [PMS] at various time intervals. The metal ions concentration range used in the kinetic studies was 2.50 × 10?5 to 1.00 × 10?4 M [Cu(II)], 2.50 × 10?4 to 2.00 × 10?3M [Ni(II)], 0.05 to 0.10 M [TAR], and µ = 0.15 M. The metal(II) tartarates, not TAR/tartarate, are oxidized by PMS. The oxidation of copper(II) tartarate at the acidic pH shows an appreciable induction period, usually 30–60 min, as in classical autocatalysis reaction. The induction period in nickel(II) tartarate is small. Analysis of the [PMS]–time profile shows that the reactions proceed through autocatalysis. In alkaline medium, the Cu(II) tartarate–PMS reaction involves autocatalysis whereas Ni(II) tartarate obeys simple first‐order kinetics with respect to [PMS]. The calculated rate constants for the initial oxidation (k1) and catalyzed oxidation (k2) at [TAR] = 0.05 M, pH 4.05, and 31°C are Cu(II) (1.00 × 10?4 M): k1 = 4.12 × 10?6 s?1, k2 = 7.76 × 10?1 M?1s?1 and Ni(II) (1.00 × 10?3 M): k1 = 5.80 × 10?5 s?1, k2 = 8.11 × 10?2 M?1 s?1. The results suggest that the initial reaction is the oxidative decarboxylation of the tartarate to an aldehyde. The aldehyde intermediate may react with the alpha hydroxyl group of the tartarate to give a hemi acetal, which may be responsible for the autocatalysis. © 2011 Wiley Periodicals, Inc. Int J Chem Kinet 43: 620–630, 2011  相似文献   

9.
Laser flash photolysis combined with competition kinetics with SCN? as the reference substance has been used to determine the rate constants of OH radicals with three fluorinated and three chlorinated ethanols in water as a function of temperature. The following Arrhenius expressions have been obtained for the reactions of OH radicals with (1) 2‐fluoroethanol, k1(T) = (5.7 ± 0.8) × 1011 exp((?2047 ± 1202)/T) M?1 s?1, (2) 2,2‐difluoroethanol, k2(T) = (4.5 ± 0.5) × 109 exp((?855 ± 796)/T) M?1 s?1, (3) 2,2,2‐trifluoroethanol, k3(T) = (2.0 ± 0.1) × 1011 exp((?2400 ± 790)/T) M?1 s?1, (4) 2‐chloroethanol, k4(T) = (3.0 ± 0.2) × 1010 exp((?1067 ± 440)/T) M?1 s?1, (5) 2, 2‐dichloroethanol, k5(T) = (2.1 ± 0.2) × 1010 exp((?1179 ± 517)/T) M?1 s?1, and (6) 2,2,2‐trichloroethanol, k6(T) = (1.6 ± 0.1) × 1010 exp((?1237 ± 550)/T) M?1 s?1. All experiments were carried out at temperatures between 288 and 328 K and at pH = 5.5–6.5. This set of compounds has been chosen for a detailed study because of their possible environmental impact as alternatives to chlorofluorocarbon and hydrogen‐containing chlorofluorocarbon compounds in the case of the fluorinated alcohols and due to the demonstrated toxicity when chlorinated alcohols are considered. The observed rate constants and derived activation energies of the reactions are correlated with the corresponding bond dissociation energy (BDE) and ionization potential (IP), where the BDEs and IPs of the chlorinated ethanols have been calculated using quantum mechanical calculations. The errors stated in this study are statistical errors for a confidence interval of 95%. © 2008 Wiley Periodicals, Inc. Int J Chem Kinet 40: 174–188, 2008  相似文献   

10.
The kinetics of electron transfer from hexacyanoferrate(II) to tris(dimethylglyoximato)-nickelate(IV), Ni(dmg)32?, to produce Fe(CN)63? and Ni(dmgH)2, follows a pseudo-first-order disappearance in the Ni(IV). The pseudo-first-order rate constants kobs are linearly dependent on [Fe(CN)64?]0 in a fiftyfold range of 2 × 10?4?1 × 10?2M, and the average values of kobs/[Fe(CN)64?]0 range from 194M?1·s?1 at pH = 5.20 to 0.2M?1·s?1 at pH = 9.07 in aqueous medium at 35°C and μ = 0.57M. Results are interpreted in terms of a probable mechanism involving rate-determining outer sphere one-electron transfer steps from the reductant and one-protonated reductant species to the unprotonated and one-protonated Ni(IV) species present in solution. The more electrophilic one-protonated reductant species apparently reacts several orders of magnitude faster than the unprotonated one.  相似文献   

11.
Mass spectrometric studies of the ions present in H2/O2/N2 flames with potassium and chlorine added have demonstrated that ionization can occur in the forward steps of K + Cl ? K+ + Cl? (II), KCl + M ? K+ + Cl? + M (IV), where M is any third body. Variations of [K+] with time in these systems have been measured and establish that the rate coefficients (in ml molecule?1 s?1) of the ion-producing steps are k2 = 5 × 10?10T?12 exp(?10 500/T) and k4 = 2.2 × 107T?3.5 × exp(?60 800/T). Coefficients for ion-ion recombination have been obtained from k2 and k4 using the equilibrium constants of (II) and (IV) and are k?2 = 1.7 × 10?9T?12 and k?4 = 1.1 × 10?17T?3, with each one in the ml molecule?1 s?1 system of units. Replacement of the N2 in one of these flames with sufficient Ar to maintain the temperature constant leaves the measured k2 and k?2 unchanged, but lowers the observed k4 and k?4. This confirms that ion-recombination in the backward step in (II) is a two-body process, whereas in (IV) it is termolecular.  相似文献   

12.
It has been shown by ESR spectroscopy that the title reaction involves abstraction of hydrogen from the phosphite, since at ?10°C the reaction has a kinetic deuterium isotope effect, kH/kD, or ~3. The rate constant for hydrogen abstraction is c. 2 × 104 M?1 s?1. There is no significant addition of alkoxyl radicals to the phosphite.  相似文献   

13.
Absolute rate constants for the reaction of O(3P) atoms with n-butane (k2) and NO(M  Ar)(k3) have been determined over the temperature range 298–439 K using a flash photolysis-NO2 chemiluminescence technique. The Arrhenius expressions obtained were k2 = 2.5 × 10?11exp[-(4170 ± 300)/RT] cm3 molecule?1 s?1, k3 = 1.46 × 10?32 exp[940 ± 200)/ RT] cm6 molecule?2 s?1, with rate constants at room temperature of k2 = (2.2 ± 0.4) × 10?14 cm3 molecule?1 s?1 and k3 = (7.04 ± 0.70)×10?32 cm6 molecule?2 s?1. These rate constants are compared and discussed with literature values.  相似文献   

14.
A combined EPR/LMR spectrometer and fast-flow system has been used to investigate the reactions HO2 + NO(k1), HO2 + OH(k2), HO2 + HO2(k3) at room temperature. The rate constants have been measured: k1 = (7.0 ± 0.6) × 10?12 cm3 s?1 (P = 7–10 Torr);k2 = (5.2 ± 1.2) × 10?11 cm3 s?1 (P = 8–10 Torr);k3 = (1.65 ± 0.3) × 10?12 cm3 s?1 (P = 2.1–24.9 Torr). The conclusion is drawn from analysis of the literature and the present work that k2 and k3 do not depend on pressure up to 1 atm.  相似文献   

15.
Three single electron charge transfer redox reactions have been studied using the faradaic rectification method. The kinetic parameters obtained for the ferricyanide-ferrocyanide redox couple are α=0.49, ka0=12×10?2 cm s?1; for the chromic-chromous system α=0.47, ka0=2×10?3 cm s?1 and for the titanic-titanous reaction α=0.49 and kao=6×10?4 cm s?1 at 27°C.  相似文献   

16.
Novel films consisting of multi-walled carbon nanotubes (MWCNTs) were fabricated by means of chemical vapor deposition with decomposition of either acetonitrile (ACN) or benzene (BZ) using ferrocene as catalyst. The electrochemical responses of MWCNT-based films towards the ferrocyanide/ferricyanide, [Fe(CN)6]3?/4? redox couple were probed by means of cyclic voltammetry and electrochemical impedance spectroscopy at 25.0?±?0.5?°C. Both MWCNT-based films exhibit Nernstian response towards [Fe(CN)6]3?/4? with some slight kinetic differences. Namely, heterogeneous electron transfer rate constants lying in ranges of 2.69?×?10?2?C1.7?×?10?3 and 9.0?×?10?3?C2.6?×?10?3?cm·s?1 were obtained at v?=?0.05?V·s?1 for MWCNTACN and MWCNTBZ, respectively. The detection limit of MWCNTACN, estimated to be about 4.70?×?10?7?mol·L?1 at v?=?0.05?V·s?1, tends to become slightly poorer with the increase of the scan rate, namely at v?=?0.10?V·s?1 the detection limit of 1.70?×?10?6?mol·L?1 was determined. Slightly poorer response ability was exhibited by MWCNTBZ; specifically the detection limits of 1.57?×?10?6 and 4.35?×?10?6?mol·L?1 were determined at v?=?0.05 and v?=?0.10?V·s?1, respectively. The sensitivities of MWCNTACN and MWCNTBZ towards [Fe(CN)6]3?/4? were determined as 1.60?×?10?7 and 1.51?×?10?7?A·L·mol?1·cm?2, respectively. The excellent electrochemical performance of MWCNTACN is attributed to the presence of incorporated nitrogen in the nanotube??s structure.  相似文献   

17.
The vibrational relaxation of pure HF(υ = 3 and υ = 4) has been studied by pumping HF directly from υ = 0 to υ = 4. The relaxation rates of υ = 3 and υ = 4 were determined to be k3T = (2.8 ± 0.4) × 10?11 cm3 molecule?1 s?1 and k4T = (7.2. ± 0.5) × 10?11 cm3 molecule?1 s?1 at 293 K. It is shown that sigle quantum energy transfer can account for all the vibrational relaxation.  相似文献   

18.
Formation of N-aminopiperidine (NAPP) in the reaction of monochloramine with piperidine was studied by varying the reagents concentrations, pH and temperature. The study was carried out in diluted solutions, recording simultaneously monochloramine concentration by UV spectrophotometry at 243 nm and hydrazine concentration at 237 nm after treatment with formaldehyde. The presence of two competitive reactions: formation of NAPP and a complex parallel reaction limiting the yield of hydrazine, was established. Reaction products were characterized by GC/MS analysis. The rate constant of NAPP formation and activation parameters were determined, k 1 = 56 × 10?3 M?1 s?1 (25°C) and k 1 = 9.3 × 106 exp(?46.5/RT) M?1 s?1, respectively.  相似文献   

19.
The rate constants of self-reactions of ketyl radicals of acetophenone in n-heptane [2k = (3.2 ± 0.5) × 109 M?1 s?1] and diphenylaminyl radicals in toluene [2k = (3.3 ± 0.5) × 107 M?1 s?1] have been determined at 298 K using the flash photolysis technique. The rate constant of ketyl radicals is equal to the calculated diffusion constant and, therefore, this reaction is diffusion-controlled. The aminyl radical recombination rate is independent of the viscosity of the toluene/vaseline oil binary mixture (0.55 ? η ? 12 cP) and this reaction is activation-controlled. Reactivity anisotropy averaging due to the cage effect has been considered for ketyl and some other radicals. On the basis of the analysis it has been proposed that ketyl recombination involves formation of not only pinacol, but also iso-pinacols.  相似文献   

20.
The interaction of cisplatin with guanine DNA bases has been investigated using ab initio Hartree–Fock (HF) and density functional levels of theory in gas phase and aqueous solution. The overall process was divided into three steps: the reaction of the monoaqua [Pt(NH3)2Cl(H2O)]+ species with guanine (G) (reaction 1), the hydrolysis process yielding the adduct [Pt(NH3)2(G) (H2O)]2+ (reaction 2) and the reaction with a second guanine leading to the product [Pt(NH3)2(G)2]2+ (reaction 3). The functionals B3LYP, BHandH, and mPW1PW91 were used in the present study, to develop an understanding of the role of the distinct models. The geometries presented for the intermediate structures were obtained by IRC calculations from the transition state structure for each reaction. The structural analysis for the intermediates and transition states showed that hydrogen bonds with the guanine O6 atom play an important role in stabilizing these species. The geometries were not very sensitive to the level of theory applied with the HF level, giving a satisfactory overall performance. However, the energy barriers and the rate constants were found to be strongly dependent on the level of calculation and basis set, with the DFT calculations giving more accurate results. For reaction 1 the rate constant calculated in aqueous solution at PCM‐BHandH/6‐311G* (k1 = 7.55 × 10?1 M?1 s?1) was in good agreement with the experiment (5.4 × 10?1 M?1 s?1). The BHandH/6‐31G* calculated gas phase rate constants for reactions 2 and 3 were: k2 = 0.9 × 10?6 M?1 s?1 and k3 = 2.99 × 10?4 M?1 s?1 in fairly good accordance with the experimental findings for reaction 2 (1.0 × 10?6 M?1 s?1) and reaction 3 (3.0 × 10?4 M?1 s?1). © 2006 Wiley Periodicals, Inc. Int J Quantum Chem, 2006  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