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 209 毫秒
1.
Densities (ρ), speeds of sound (u), isentropic compressibilities (ks), refractive indices (nD), and surface tensions (σ) of binary mixtures of methyl salicylate (MSL) with 1-pentanol (PEN) have been measured over the entire composition range at the temperatures of 278.15 K, 288.15 K, and 303.15 K. The excess molar volumes (VE), excess surface tensions (σE), deviations in speed of sound (Δu), deviations in isentropic compressibility (Δks), and deviations in molar refraction (ΔR) have been calculated. The excess thermodynamic properties VE, σE, Δu, Δks, and ΔR were fitted to the Redlich–Kister polynomial equation and the Ak coefficients as well as the standard deviations (d) between the calculated and experimental values have been derived. The surface tension (σ) values have been further used for the calculation of the surface entropy (SS) and the surface enthalpy (HS) per unit surface area. The lyophobicity (β) and the surface mole fraction (x2S) of the surfactant component PEN have been also derived using the extended Langmuir model. The results provide information on the molecular interactions between the unlike molecules that take place at the surface and the bulk.  相似文献   

2.
The faradaic impedance of the surface redox system benzo(c)cinnoline-dihydrobenzo(c)-cinnoline is studied experimentally in aqueous medium between pH 5 and 13. The variations of the impedance components are in good accord with the theoretical predictions. A V-shaped curve is found for log ks=f(pH) (ks=rate constant of the surface electrochemical reaction). It is estimated that the determination of rate constant values up to 2×104 s?1 on a mercury electrode is possible by this method.  相似文献   

3.
Standard electrochemical rate constants kexob have been measured for the two sequential one-electron oxidations of triply-bonded dirhenium(II) complexes Re2X4(PR3)4, where X = Cl or Br, and PR3 = a monodentate tertiary phosphine, and also for allied complexes in methylene chloride, acetonitrile, and N,N-dimethylformamide at platinum electrodes. The objective is to explore the possible dependence of kexob upon the known differences in the structural changes accompanying electron transfer. Although significant and even substantial variations in kexob were observed for closely related redox couples, the reactivity trends appear to be chiefly a consequence of differences in electrostatic work terms at the metal/solution interface. Comparisons are made with recent results for mononuclear organometallic redox couples.  相似文献   

4.
In order to quantify our fluorogenic molecular probe studies on the radiation-induced polymerization of methyl methacrylate (MMA), measurements have been made of the monomer conversion, C M, and polymer molecular weight distribution as a function of dose for bulk MMA polymerized by steady-state (60Co γ-rays) and nanosecond-pulsed (3 MeV electrons) radiation. In all cases, C M was found to increase close to linearly with dose up to ca. 30%. Above this conversion, autoacceleration of polymerization (the gel or Trommsdorff effect) occurs in the γ-irradiated samples. From the low-conversion steady-state data, using dose-rates of 0.21 and 2.7 Gy/s, the parameter 〈k p2 G(R.)/2〈k t〉 = 0.011 and 0.015 lmol-1 s-1 (100 eV)-1 respectively have been determined (with 〈k p〉 and 〈k t〉 the average rate coefficients for bimolecular propagation and termination respectively, and G(R.) the yield of the chain-initiating free radicals per 100 eV (G = 1 (100 eV)-1 0.1036 μ mol J-1)). From the 5 Hz repetitive pulse data the value of 〈k pG(R.) = 1700 lmol-1 s-1 (100 eV)-1 has been determined. Taking 〈k p〉 = 342 lmol-1 s-1 from the literature results in G(R.) = 5.0 (100 eV)-1 and 〈k t〉 = 2.7 × 107 lmol-1 s-1.  相似文献   

5.
The influence of polyvinyl alcohol on the kinetics of cadmium ion reduction in 1M aqueous KCl solution at the dropping mercury electrode was investigated using the impedance method. It was found that small additions of PVA did not change the diffusion coefficient much, but affected only the rate of charge transfer process. The electroreduction of cadmium in the presence of PVA became less reversible. The standard rate constant dependence on coverage degree is described by the following equation: ks=ks0(1?ν)n, where n=2.  相似文献   

6.
A measurement approach is described and data are presented which demonstrate the ability to effect a.c. cyclic voltammetric measurements with the on-line digital FFT approach to faradaic admittance data acquisition. The equipment utilized enables complete faradaic admittance spectra to be obtained at an effective spectrum acquisition rate of 10 s?1, so that the d.c. potential range encompassed by a typical cyclic wave can be encompassed with adequate resolution in the Edc dimension in ≥6 s, approximately. The instrument features dynamic, computerized measurement and compensation of the non-faradaic ohmic resistance and double-layer capacitance contributions to the acquired total cell admittance. Measurements with quasi-reversible systems yield the expected faradaic admittance and phase angle responses over a quite generous bandwidth. Applications to mercury and platinum electrodes are illustrated.  相似文献   

7.
Convolution voltammetry was used to evaluate the rates of heterogeneous charge transfer to ferrocene groups in poly(vinylferrocene) and to Ru(bpy)2+3 in Nafion-modified electrodes under semi-infinite conditions. This technique allows correction for uncompensated resistance and double layer capacitance, as well as detrmination of the diffusion coefficient, D, transfer coefficient, α, and half-wave potential, E1/2, from a single cyclic voltammogram. Vinylferrocene in solution and a bound copolymer of vinylferrocene and styrene in a ratio of 58:42 were also examined. For the polymer films, the heterogeneous charge transfer rate constants, k°, are 10?4k° ≥ 10?5 cm/s; these values are about two order of magnitude smaller than those for the similar species in homogeneous solution. The values of k°/D1/2, however, are comparable to those in soluton; 10 > (k°/D1/2) > 0.1 s?1/2.  相似文献   

8.
The temperature dependence of the fluorescence and fluorescence excitation spectra of all-trans diphenyl hexathene (DPH) and octatetraene (DPO) in six solvents confirms the S1(1Ag*) and S2(1Bu*) state assignment, and determines their energy difference ΔE. The S1 fluorescence rate parameter kF depends on ΔE, the solvent refractive index n, the S2 (n = 1) fluorescence rate parameter kF20 (2.23 × 108 s?1 for DPH, 2.33 × 108 s?1 for DPO), and the S2-S1 coupling matrix element V (745 cm?1 for DPH, 500 cm?1 for DPO). The S1 fluorescence is induced by 1Bu*-1Ag* potential interaction (PI), via a bu vibrational mode (≈ 900 cm?1), and not by vibronic coupling. The main S1 radiationless transition, rate parameter kR, is thermally-activated internal rotation through an angle θ about the central ethylenic bond(s). The PI distorts the S1 (θ) potential surface and thus influences kR.  相似文献   

9.
Uncompensated resistance (Ru) has a distoring effect on normalized potential sweep voltammetry (NPSV) slopes. This provides a simple and effective method to determine Rfo, the value of the potentiostat feedback resistance necessary for full compensation. If the NPSV range is divided into overlapping segment, 1 and 2, corresponding to IN of 0.20–0.50 and 0.50–0.80 respectively, the slopes m1 and m2 differ significantly when Rf differs form Rfo. The difference, m1-m2, is negative for Rf<Rfo and positive for Rf>Rfo. Fine tuning of the potentiostat Rf setting so that the average value of the difference is the theoretical value can be accomplished in a minimum of time. Under these conditions, m1 and m2, as well as mT, the slope of the entire correlation have very nearly the same values. Linear equations were derived from theoretical data which allow heterogeneous rate constants to be obtained directly from NPSV slopes. The precision in the NPSV slopes was observed to be of the order of ±0.002 which implies that the method should give reliable rate constants as great as 10cm s? at a voltage sweep rate of 100V s?1. The method is demonstrated using the reduction of benzonitrile and perylene in N,N-dimethylformamide and acetonitrile as examples.  相似文献   

10.
Based on the table of the effective ionic radii of R. D. Shannon (Acta Crystallogr. Sect. A32, 751, 1976), the linear relation of the cationic radii on the coordination numberk has been observed:
rk=r0+dk?0.0236kz,
wherer0 is the radius of free cation, z its valence, andd = 0.1177 - 0.0081 z - 0.0347 r0 - 0.0050 zr0. The analogous relation for O2? anion has been found to ber′k′ = r′0 + a′k′ = 1.328 + 0.0118 k′. By addingrk andr′k′ corresponding to the same bond strength new dependence between bond lengthR and bond strengths has been established for cation-oxygen bonds:
s=dzR?R0,
whereR0 = r0 + r′0. The necessity to choose a standard state for ionic radii and for bond strength is pointed out and argued. Structures of simple oxides at room temperature and at normal pressure are chosen as standard state at which the sum of the strengths of bonds around the cation is assumed to be exactly equal to itsz.Standard radii of free ions r0 are determined for about 230 ions and are listed. As in Shannon's table some ionic radii were found to be negative. Physical sense can be hardly attributed to this finding. To avoid the negative radii a new scale ofabsolute ionic radiiρ0 is proposed, based on assumption thatρ0(H1+) = 0 instead ofr0(H1+) = ?h = ?0.499 (the size of the proton is known to be of the order of 10?5A?, i.e., much less than the accuracy of determination of ionic radii). Consequentlyρ0 for all cations is assumed to beρ0 = r0 + h and for anionsρ′0 = r′0 ? h (e.g.,ρ′0(O2?) = 0.829A?). Bond-length-bond-strength relationship expressed by the equation indicated above is rationalized in terms of the new proposedelectrostatic hover model of crystal structure. In this model ions are of constant size (r0 orρ0), they do not touch each other, but they are maintained at distancesL = R - R0 by the electrostatic forces. This remains in agreement with the suggested (J. Zio´?kowski,J. Catal.84, 317, (1983)) linear relation between bond strength and bond energyE = Js which will be proved in the forthcoming paper (J. Zio´?kowski and L. Dziembaj,J. Solid State Chem.57, 291 (1985)).  相似文献   

11.
In this paper, excess thermodynamic functions have been computed from the measured values of density, viscosity, and refractive index at T = (298.15, 303.15, and 308.15) K, ultrasonic velocity at T = 298.15 K over the entire mixture composition range of (anisole with ethanol, propan-1-ol, propan-2-ol, butan-1-ol, pentan-1-ol, or 3-methyl butan-1-ol). Excess molar volume, VE has been calculated from densities, whereas deviations in viscosity, Δη, were computed from the measured viscosities. From ultrasonic velocities, isentropic compressibilities were calculated, from which deviations in isentropic compressibility, Δks have been computed. Lorenz-Lorentz mixture rule was used to compute molar refractivity, R from refractivity index data and from these data, deviations in molar refractivity, ΔR have been computed. Computed thermodynamic quantities have been fitted to Redlich and Kister polynomial equation to derive the coefficients and standard errors between experimental and predicted quantities. Intermolecular interactions between anisole and alkanols have been studied based on the computed excess thermodynamic quantities.  相似文献   

12.
13.
The triplet-triplet energy transfer from benzaldehyde to biacetyl and the competing self-quenching between triplets and ground state molecules of benzaldehyde were investigated in the dilute vapor phase by monitoring the phosphorescence (T1(nπ*)So) decay of benzaldehyde. Following excitation into the S1(nπ*)S0 absorption band, a triplet self-quenching rate constant of kSQ=(2.4±0.1) × 104 s?1 Torr?1, corresponding to a gas-kinetic cross section of σSQ=0.22 A2, was measured. The collision-free lifetime of the benzaldehyde triplet was found to be 2.3 ± 0.4 ms. Substitution of the aldehydic proton by deuterium reduces kSQ by a factor of two: complete deuteration of the molecule has no further effect. Under the same excitation conditions, the energy transfer rate to biacetyl is kET=(2.8 ± 0.1) × 106 s?1 Torr?1, with σET = 24 A2. This process is not influenced by deuteration.  相似文献   

14.
Cavity ring‐down UV absorption spectroscopy was used to study the kinetics of the recombination reaction of FCO radicals and the reactions with O2 and NO in 4.0–15.5 Torr total pressure of N2 diluent at 295 K. k(FCO + FCO) is (1.8 ± 0.3) × 10−11 cm3 molecule−1 s−1. The pressure dependence of the reactions with O2 and NO in air at 295 K is described using a broadening factor of Fc = 0.6 and the following low (k0) and high (k) pressure limit rate constants: k0(FCO + O2) = (8.6 ± 0.4) × 10−31 cm6 molecule−1 s−1, k(FCO + O2) = (1.2 ± 0.2) × 10−12 cm3 molecule−1 s−1, k0(FCO + NO) = (2.4 ± 0.2) × 10−30 cm6 molecule−1 s−1, and k (FCO + NO) = (1.0 ± 0.2) × 10−12 cm3 molecule−1 s−1. The uncertainties are two standard deviations. © 2001 John Wiley & Sons, Inc. Int J Chem Kinet 33: 130–135, 2001  相似文献   

15.
The electrochemical properties of FAD adsorbed on graphite were studied with cyclic voltammetry. A film with a surface coverage of up to 2×10?8 mol cm?2 was formed after 4 h. The film-covered electrodes were stable for long times and they were investigated in buffers without any FAD. A single well-behaved wave was observed, k0≌1 s?1. The pH dependence of E1/2 was studied. On glassy carbon, platinum and gold much less FAD was adsorbed and it was removed with one rinse.  相似文献   

16.
Reactivities of free radical oxidants, .OH, Br-·2 and Cl3COO. and a reductant, CO-·2, with trypsin and reactive protein components were determined by pulse radiolysis of aqueous solutions at pH 7, 20°C. Highly reactive free radicals, .OH, Br-·2 and CO-·2, react with trypsin at diffusion controlled rates, k(.OH + trypsin) = 8.2 × 1010 M-1 s-1, k(Br-·2 + trypsin) = 2.55 × 109 M-1 s-1 and k(CO-·2 + trypsin) = 2.6 × 109 M-1 s-1. Moderately reactive trichloroperoxy radical, k(Cl3COO. + trypsin) = 3 × 108 M-1 s-1, preferentially oxidizes histidine residues. The efficiency of inactivation of trypsin by free radicals is inversely proportional to their reactivity. The yields of inactivation of trypsin by .OH, Br-·2 and CO-·2 are low, G(inactivation) = 0.6-0.8, which corresponds to ∾ 10% of the initially produced radicals. In contrast, Cl3COO. inactivates trypsin with ∾ 50% efficiency, i.e. G(inactivation) = 3.2.  相似文献   

17.
2‐Phenylethanol, racemic 1‐phenyl‐2‐propanol, and 2‐methyl‐1‐phenyl‐2‐propanol have been pyrolyzed in a static system over the temperature range 449.3–490.6°C and pressure range 65–198 torr. The decomposition reactions of these alcohols in seasoned vessels are homogeneous, unimolecular, and follow a first‐order rate law. The Arrhenius equations for the overall decomposition and partial rates of products formation were found as follows: for 2‐phenylethanol, overall rate log k1(s−1)=12.43−228.1 kJ mol−1 (2.303 RT)−1, toluene formation log k1(s−1)=12.97−249.2 kJ mol−1 (2.303 RT)−1, styrene formation log k1(s−1)=12.40−229.2 kJ mol−1(2.303 RT)−1, ethylbenzene formation log k1(s−1)=12.96−253.2 kJ mol−1(2.303 RT)−1; for 1‐phenyl‐2‐propanol, overall rate log k1(s−1)=13.03−233.5 kJ mol−1(2.303 RT)−1, toluene formation log k1(s−1)=13.04−240.1 kJ mol−1(2.303 RT)−1, unsaturated hydrocarbons+indene formation log k1(s−1)=12.19−224.3 kJ mol−1(2.303 RT)−1; for 2‐methyl‐1‐phenyl‐2‐propanol, overall rate log k1(s−1)=12.68−222.1 kJ mol−1(2.303 RT)−1, toluene formation log k1(s−1)=12.65−222.9 kJ mol−1(2.303 RT)−1, phenylpropenes formation log k1(s−1)=12.27−226.2 kJ mol−1(2.303 RT)−1. The overall decomposition rates of the 2‐hydroxyalkylbenzenes show a small but significant increase from primary to tertiary alcohol reactant. Two competitive eliminations are shown by each of the substrates: the dehydration process tends to decrease in relative importance from the primary to the tertiary alcohol substrate, while toluene formation increases. © 1999 John Wiley & Sons, Inc. Int J Chem Kinet 31: 401–407, 1999  相似文献   

18.
A longlived triplet state is shown by 119Sn CIDNP for distannanes R6Sn2 upon irradiation (ET = 300 kJ), which may be split into 2 radicals R3Sn·(ka), but also into R5Sn2· + R· (kb. With R = Me, ka is more important, but with R = Et, Bu, kakb. Compounds R6Sn2 are cleaved by R· into R4Sn + R3Sn·, the effectivity being increased in the series t-Bu < i-Pr << Pr < Et < Me ≈ Ph. For Me6Sn2 and Me·, k = 0.85 X 104 l mol?1 sec?1 at 23°C. The distannane Et5Sn2Br is attacked by Me· predominantly, by Pr· and Bu· exclusively at the halogenated Sn atom. Diacetyl peroxide reacts quickly with R6Sn2 via a nonradical mechanism yielding MeCOOSnR3, as does dibenzoyl peroxide analogously but more slowly. In both cases a free-radical reaction is superposed (13C CIDNP) giving R4Sn, CO2, and RCOOSnR3. It is stopped by t-BuBr, but increased markedly at higher temperatures.  相似文献   

19.
In the flash photolysis of SiBr4 both the absorption and the emission spectra corresponding to the B̃2Σ−X̃2Π transition of SiBr have been observed. A broad, structureless absorption band has also been detected in the 340–400 nm region which could be assigned to the hitherto unreported à 1B1−x̃ 1A1 transition of SiBr2. The decay of both absorption spectra followed first-order kinetics yielding the pseudo-first-order rate constants: k(SiBr)=2.6 × 104s−1 and k(SiBr2) = 8.9 × 103−1. Assuming that the principal reactions consuming these intermediates are SiBr+SiBr4→Si2Br5 and SiBr2+SiBr4→ Si2Br6, the second-order rate constants have the values k(SiBr)= 9.7×109 M−1s−1 and k(SiBr2)= 3.3×108M−1s−1.  相似文献   

20.
The voltammetric curves for a 2-step sequential reaction involving a 2nd order adsorption process (such as the catalytic mechanism met in the anodic oxidation of hydrogen) are investigated theoretically by computer calculations. Steady state cyclic voltammetry (s.s.c.v.) is particularly treated by repeating forward and backward sweeps as many times as necessary to reach the steady state. For both l.p.s.v. and s.s.c.v., the usual characteristic features are considered, i.e. EM, θM, iM as a function of sweep rate υ, especially in the range of υ, where the computed data fit the results derived from analytical solution (quasi-reversible charge transfer, and irreversible charge transfer). In s.s.c.v., indications are given to obtain the kinetic parameters (k1, k2, α) from experimental plots.  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