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.
The Arrhenius plot (logarithmic plot vs. inverse temperature) is represented by a straight line if the Arrhenius equation holds. A curved Arrhenius plot (mostly concave) is usually described phenomenologically, often using polynomials of T or 1/T. Many modifications of the Arrhenius equation based on different models have also been published, which fit the experimental data better or worse. This paper proposes two solutions for the concave-curved Arrhenius plot. The first is based on consecutive A→B→C reaction with rate constants k1k2 at higher temperatures and k1k2 (or at least k1 > k2) at lower temperatures. The second is based on the substitution of the temperature T the by temperature difference TT0 in the Arrhenius equation, where T0 is the maximum temperature at which the Arrheniusprocess under study does not yet occur.  相似文献   

2.
The reductions of several substituted acetophenones using supercritical 2-propanol were carried out to estimate the Hammett's reaction constant (ρ=0.33). Also, the reduction of acetophenone using supercritical deuteriated 2-propanol was carried out to determine the rate-determining step. The kinetic isotope effects were observed in the reduction using 2-deuterio-2-propanol (kH/kD=1.6) and O-deuterio-2-propanol (kH/kD=2.0). These findings suggest that the reaction proceeds via a cyclic transition state between acetophenone and 2-propanol similar to that of the Meerwein-Ponndorf-Verley reduction.  相似文献   

3.
Selective extraction of antioxidants with molecularly imprinted polymers   总被引:3,自引:0,他引:3  
Molecular imprinting technology can be used to generate specific artificial polymeric receptors, i.e., high affinity stationary phases, as already shown for peptides and many other food ingredients. In this work, polymers have been molecularly imprinted with three exemplary antioxidants, butylated hydroxyanisole (BHA), butylated hydroxytoluene (BHT) and propyl gallate (PG). Results of chromatographic evaluations are presented, demonstrating the specificity of the imprinted polymers (MIP) compared to non-specific control polymers (CP). For a BHA imprinted polymer and the respective analyte BHA a separation factor of α=kBHA-MIP′/kCP′=1.12 was found, whereas for the BHT-MIP a higher selectivity was determined with α=kBHT-MIP′/kCP′=1.47. Using the PG selective MIP for PG as analyte a value of α=kPG-MIP′/kCP′=1.24 could be achieved. Furthermore, results of extraction procedures based on MIP and CP phases are presented, demonstrating for instance a higher selectivity of the BHA-MIP in comparison with its CP when using methanol or acetonitrile as analyte solvent, and of the PG-MIP when performing the extraction in acetonitrile.  相似文献   

4.
A novel approach is demonstrated for measuring rates of the consecutive acid-base catalyzed hydrolysis reactions of (3-glycidoxypropyl)trimethoxysilane (GPTMS) and (3-aminopropyl)triethoxysilane (APTES) in dilute aqueous solution using liquid chromatography with inductively coupled plasma atomic emission spectrometric (ICP-AES) detection. The hydrolysis reactions are monitored by sampling kinetic solutions in a timewise manner and performing liquid chromatographic separations of the parent silane and organosilicon hydrolysis products. The column effluent is fed into the ICP through a direct injection nebulizer for online monitoring of silicon atomic emission at 251.611 nm, producing a series of silicon chromatograms for each kinetic run. Reversed phase separations are effected using acetonitrile-water gradients and are complete in 6 min or less. The systematic changes in peak areas provide information from which the rate constants of the consecutive hydrolysis reactions (k1, k2, and k3) are obtained by non-linear regression modeling. Using a quenching scheme, hydrolysis half-lives as brief as 3 min for the parent silane can be monitored. For each compound, a series of rate constants are obtained over a range of pH and buffer concentration, permitting estimation of the catalytic constants kH3O+ and kOH for the consecutive acid-base catalyzed hydrolysis reactions by multiple regression analysis.  相似文献   

5.
We have derived an equation for the instantaneous limiting current in pulse polarography with an ECE mechanism. That equation has been derived for the expanding sphere electrode model (and also for the expanding plane electrode model) with the condition (k1+k2)>>1, k1 and k2 being the rate constants of the chemical reaction. We show that by an adequate choice of the time of application of the potential and time of the drop growth prior to potential application, it is possible — if the equilibrium constant for chemical reaction is known — to widen the interval of values for k1 and k2 that may be determined using this technique. At the same time, we also show that if the electrode sphericity is not taken into account, the values obtained for k1 and k2 are always lower than the real ones.  相似文献   

6.
Many overview papers have been published with recommended nuclear data for use in the k 0 method of NAA and made available in scientific journals or in the form of a downloadable database. In September 2009, the k 0-International Scientific Committee formed the k 0-Nuclear Data Committee (k 0-NDC) whose first task was to collect all these data at a single place to facilitate updating and to correct any evident errors. This task of the k 0-NDC was successfully completed in March 2012 when the 2012 recommended k 0 database was published in the form of an Excel file.  相似文献   

7.
Photoirradiation of dipyridamole (DIP) solution at λ ⪢ 390 nm leads to a reduction of the intensity of the absorption band without change in its appearance. This reduction is due to the breaking of the π-conjugation chain of the DIP molecule. The experimentally measured rate constant k is proportional to the concentration of oxygen and to the ratio of the rates of photoreaction, kr, and of radiation, ko. In homogeneous solutions the values of kr and ko are three-times greater at pH 5.0 than at pH 7.0. So, the photoreaction is more effective when the DIP molecules are completely protonated. In alcoholic and alkaline solutions (pH 13.5), the deprotonation of DIP molecules is responsible for a significant reduction of kr. In the presence of microheterogeneous systems, lysophosphatidylcholine (L-PC) micelles and bovine serum albumin (BSA), a significant reduction of kr is observed. The value of kr in this case depends on the pH of the solution as well as on the concentration of L-PC and BSA. In the presence of L-PC or BSA, different values of k could be associated to free and bound DIP molecules. A kinetic procedure is proposed which permits the evaluation of the binding constants as well as the kinetic constants. The binding constant of DIP to L-PC micelles is estimated as 1.28 × 104 M−1 and the value of the limiting effective rate constant k is similar to the value obtained in ethanol. This is evidence for the localization of DIP molecules in the polar region of L-PC micelles. The value of k in this case is three-times greater at pH 5.0 than at pH 7.0 as occurs in homogeneous solutions. The binding to BSA was also studied and binding constants of (1.8±0.2) × 104 M−1 and (7.8±0.9) × 104 M−1 were obtained at pH 7.0 and pH 5.0, respectively. In this case the ratio k(ph 5.0)/k(ph 7.0) is also equal to three, the same as in homogeneous solutions or in the presence of L-PC micelles. This implies that DIP molecules maintain their protonation form in the presence of microheterogeneous systems.  相似文献   

8.
Rate constants for vibrational energy transfer have been measured in the system N2(v= 1) +CO(v=0) as a function of energy mismatch by using isotopic derivatives of N2 and of CO. These rate constants have been determined both in the gas phase and in liquid argon solution at 85 K. For non-resonant processes we find kL=kG for the nearest to resonant system we find kL <kG. This result is considered to arise because long range forces important in near-resonant energy transfer operate to a lesser extent in the liquid phase.  相似文献   

9.
《Chemical physics letters》1985,117(2):132-137
Absolute rate constants, k, of the reaction OH + HNO3 were determined using a pulsed laser photolysis-resonance absorption technique. The measured values, in cm3 mol-1 s-1 at ±3σ, 10-10k(1–16 Torr HNO3) = 7.57 ± 0.64, k(500 Torr N2) = 7.20 ± 0.66 and k(600 Torr SF6) = 8.37 ± 0.45, indicate that any pressure effect on k at 297 K is less than the experimental uncertainty of 10%.  相似文献   

10.
The reversal in coupling (from antiferromagnetic to ferromagnetic order with increasing x) in EuxLa1−xAl2 ternaries is successfully analyzed in terms of the RKKY formalism. The Fermi wavevector k required to account for the behavior of these ternaries is k = 0.94 k0, which is in close agreement with the k-value previously obtained for GdAl2 and GdAl2-based ternaries from NMR work. The similarity of behavior of Gd0.1La0.9Al2 and Eu0.1La0.9Al2 and their observed Curie temperatures can also be understood using the same theoretical framework and approximately the same k-values.  相似文献   

11.
It is shown that the eigenvalues Ei of a Hermitian matrix H with matrix elements Hij = ΣkAkijak, where Akij are known numbers and ak a set of parameters, can be exactly expanded as Ei = Σk(?Ei?ak)ak. This property is applied to the analysis of the optical spectra of transition metal ions in crystals proposed by L. Pueyo, M. Bermejo, and J. W. Richardson (J. Solid State Chem.31, 217, 1980), and it is shown that this method represents the best fit of the Hamiltonian eigenvalues to the observed (or calculated) spectrum. Further advantages of using this property, in connection with the spectral analysis, are the minimization of the errors associated with the numerical approximations and a reduction in computer time. In the molecular orbital calculation of the optical or uv spectra of these systems, this linear expansion of the eigenvalues give a detailed interpretation of the improvements produced by refined calculations, such as those including configuration interaction. In particular, the changes in one-electron energy and in open-shell repulsion interactions associated with the refinement can be clearly and easily formulated. As examples, the computed spectra of CrF4?6 and CrF3?6 are discussed.  相似文献   

12.
At constant temperature (isothermal) maintained throughout in the capacitive discharge technique, the measured absorbance at any time t due to concentration of analyte atoms can be given by: absorbance = p[A]0{k1/(k1?k2)}[exp(?k2t)-exp(?k1t)], where p is a function of the oscillator strength (a constant) and the efficiency with which the analyte atoms are produced, [A]0 is the initial concentration of the analyte atoms, k1 and k2 are first-order rate constants for formation and decay of analyte atoms, respectively. This technique yields k1?k2 and k1t?k2t; and so the above equation reduces to: absorbance ?p[A]0, resulting in large enhancement in sensitivity. In the case of lead, the immediate precursor of the gaseous lead monomer is the gaseous lead dimer, which is partly lost by diffusion of the lead dimer with a first-order rate constant, k3. The kinetic parameters k1, k2 and k3 have been evaluated, and the values of k1 at different temperatures used to draw the Arrhenius plots, from which activation energies of the rate-determining steps have been determined. The activation energies have been used to elucidate atomization mechanisms by extensive correlation of the experimental energy values with the literature values.  相似文献   

13.
The kinetics of OH(ν = 0) and OH(ν = 1) have been followed using pulsed photolysis of H2O or HNO3 to generate hydroxyl radicals, and time-resolved, laser-induced fluorescence to observe the rates of their subsequent removal in the presence of HCl or HBr. The experiments yield the following rate constants (cm3 molecule?1 s?1) at 298 ± 4 K: OH(ν = 0) + HCl: ko = (6.8 ± 0.25) × 10?13; OH(ν = 0) + HBr: ko = (11.2 ± 0.45) × 10?12; OH(ν = 1) + HCl: k1 = (9.7 ± 1.0) × 10?13; OH(gn = 1) + HBr; k1 = (8.1 ± 1.05) × 10?12 For OH(ν = 1), the measurements do not distinguish between loss by reaction and relaxation, and the fact that k1 > ko for HCl is tentatively attributed to relaxation, probably by near-resonant vibrational—vibrational energy transfer. Clearly, neither of these exothermic, low-activation-energy reactions is enhanced to any great extent, if at all, by vibrational excitation of the OH radical.ft]*|Present address: Battelle/Pacific Northwest Laboratories, P.O. Box 999, Richland, Washington 99352, USA.  相似文献   

14.
The dependence of product distribution from the various experimental parameters (rates, concentration diffusion layer thickness, current density) is analyzed for a reaction scheme involving the competition between dimerization and first-order deactivation of the intermediate resulting from the first electron uptake (or removal). Two cases are considered, corresponding to the product of the first-order deactivation being either electroinactive or undergoing a further electron transfer at the electrode or in the solution. It is shown that the dimer yield is not very sensitive to the further fate of the deactivation product. It is in all cases a function of the competition parameter σ=k1k2?3/2D1/2 c0 δ?1 in potentiostatic conditions and σΔ=k1k2?3/2D?1/2(i0/FS) in galvanostatic conditions (k1, k2 rate constants of the dimerization and deactivation reactions respectively; D diffusion coefficient; c0 concentration of substrate; δ diffusion layer thickness; (i0/S) current density). The reduction of CO2 in DMF with competitive formation of oxalate and formate is taken as an example illustrating the theoretical analysis.  相似文献   

15.
The standard rate constants (k s) of charge transfer on a glass carbon electrode were determined for the Cr(III)/Cr(II) redox pair in the NaCl-KCl-K3CrF6, KCl-K3CrF6, and CsCl-K3CrF6 systems at 973–1173 K by cyclic voltammetry. The k s constant was found to increase at elevated temperatures and the following nonmonotonic dependence of k s on the nature of the outer-spheric cation was found: k s (CsCl) > k s (NaCl-KCl) > k s (KCl). On the basis of quantum-chemical data for the M3CrF6 + 18MCl (M = Na, K) model systems, it was shown that the complex chromium particles with four or five outer-spheric sodium or potassium cations had maximum thermodynamic stability. Quantum-chemical calculations were performed to interpret the experimental data on the effect of the second coordination sphere of the complexes on the standard charge transfer rate constants.  相似文献   

16.
The excimer lifetime τD and the excimer fluorescence efficiency for 1-methylnaphthalene in ethanol have been determined between −30 and 60°C. Expressing the rate constant for excimer deactivation. kD, in terms of radiative kFD) and nonradiative (kID) processes as kD = kFD + kID it is found that kFD is independent of tempeature and kID = 5 × 106 + 3.2 × 1011 exp (-ΔE/RT) sec−1, where ΔE = 6.7 kcal mole−1. The behaviour of kFD and kID with temperature and the appearance of isoemissive points in a limited region of temperature are discussed.  相似文献   

17.
The electron self-exchange rates (kex) of viologen and its derivatives are estimated by using microelectrode voltammetry in poly(ethylene glycol) films. The dependences of supporting electrolyte concentration and sizes of viologen and its derivatives on kex and diffusion coefficients (D) are discussed. Results show that kex increases with the decrease of supporting electrolyte concentration and sizes of reactants.  相似文献   

18.
The kinetics and stoichiometry of the formation of active oxygen (AO) in the acidic decomposition of trimeric (TATP) and dimeric (DADP) cyclic acetone peroxides are considered. Fe(III) produced as a result of Fe(II) oxidation with active oxygen has been determined using rhodanide procedure. The kinetics of the formation of active oxygen is described by a first order equation. The effective rate constant of TATP decomposition depends on the Hammett acidity function H 0: log k eff = ?H 0 ? 2.6 (k eff is in s?1). Consequently, the decomposition rate of TATP is limited by protonation. In the HCl and H2SO4 concentration range from 0.006 to 2.9 mol/L, the decomposition of DADP occurs with k eff = 0.0010 ± 0003 s?1 at a Fe(II) concentration of 3.5 mmol/L and k eff depends linearly on the concentration of Fe(II).  相似文献   

19.
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   

20.
The response of an amperometric oxygen electrode is studied theoretically and experimentally for the case of significant liquid film resistance at the outer side of the membrane. The behaviour of the probe in a gas stream is predicted by using a transfer function which involves an electrode model including oxygen diffusion within the electrode. The effects of the liquid film which exists at the membrane surface when dissolved oxygen concentrations are measured, are taken into account by modifying the transfer function. The new expression obtained is used to model the step response of the probe in a 2-1 stirred tank fermentor filled with water or xanthane solutions at different concentrations. First, the results are used to correct automatically the steady-state voltage readings of the probe. Secondly, the probe transfer function is used to evaluate k1a by dynamic measurements: the response to a step change in the gas concentration is transformed via the fast Fourier transform algorithm and k1a is identified in the Fourier domain by a Gauss-Newton algorithm. Data acquisition, Fourier transform and k1a identification are implemented on-line on a HP-87 computer. This method of obtaining k1a values appears to be a generalized moment method. It is shown that it is necessary to consider the liquid film dynamics around the probe in the actual fermenting conditions to evaluate k1a successfully.  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