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.
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   

2.
Spectrophotometric methods were utilized for stability constant determinations of the Fe(III) interaction with glycinehydroxamic acid (GX) at I = 0.15 M NACl and T = 25°C. Program SQUAD II was used to assess the absorbance data in the wavelength range 300–520 nm. Four constants were determined for 1:1:1, 1:1:0, 2:1:1 or 3:1:3 and 2:1:0 complex species in the pH range 1.0–7.5. The kinetics of the interactions of Fe(III) with GX were also studied in the pH range 1.0–3.0 by the stopped flow method. The observed rate constant at a given pH was kobs = A + BTGX. The parameters A and B are functions of pH in the range 1.7–3.0 and only A is a function of pH in the range 1.0–1.7. The mechanism of complex formation was discussed in the light of the experimental results and the equilibrium study. It has been concluded that FeOH2+ is the reactive species in the complex formation of FeGXH3+ species while Fe(OH)2+ is the reactive species in the complex formation of FeGX2+ species.  相似文献   

3.
The simultaneous determination of iron(III) and titanium(IV) with diantipyrinylmethane (DAPM) based on dual-wavelength spectrophotometry is described. The absorbances at 388 nm, 470, and 514.9 (A388, A470, A514.9, respectively) are measured and a ratio k (= A388/A514.9) of 3.64 is introduced to allow simultaneous determinations of iron and titanium. The apparent molar absorptivities obtained by using the differences in absorbance, A388—A514.9, for titanium and A470 × k — A388 for iron, are 1.41 × 104 and 1.13 × 104 1 mol?1 cm?1, respectively. The calibration graphs are linear up to 20 mg 1?1 iron(III) oxide and 5 mg 1?1 titanium(IV) oxide. The proposed method was applied successfully to the determination of iron and titanium in silicate rocks. The protonation equilibria of DAPM were also studied; Ka1 and Ka2 are estimated as 101.10 and 100.75, respectively.  相似文献   

4.
In order to analyse spectrometric and potentiometric data for a multicomponent system we have derived a relationship A(Uh) relating directly absorbance and redox potential changes. The advantages of such a representation A(Uh) are discussed with regard to the usual one [log(ox/red), Uh] and compared to other methods.The theoretical limits of the resolution of the midpoint potentials of two components are determined in the case of one and two-electron processes. These limits, which are function of the physical parameters and of the experimental precisions, can be less than 5 mV. The limit for the detection of a minor absorbing species is also discussed.The analysis technique was tested experimentally with mixtures of dyes. In the following paper of this issue the potentiometric-resolution of the c type cytochromes in yeast mitochondria is presented.  相似文献   

5.
When data on the variation with time of the absorbance of a reactant or product are used to evaluate the rate constant of a first- or pseudo-first-order reaction, the precision of the result depends on the precisions with which both the time and the absorbance are measured. The natures of the dependences, and the ways in which they are affected by both constant and linearly varying background absorbances, are examined. If the standard error σt of a measurement of time is below about 0.005 t12, the standard error of the rate constant is virtually identical for an experiment in which the concentration of the reactant is followed as for one in which the concentration of the product is followed, but for larger values of σt it is better to follow the concentration of the reactant. In any event, errors in the measurements of time are much more likely to be significant than they are usually assumed to be. Other sources of error in such experiments, including constant and time-dependent background absorbances, are examined more briefly, with emphasis on the requirements that should be satisfied in work of the highest quality.  相似文献   

6.
《印度化学会志》2021,98(12):100232
The current research focuses on the computation of absorption spectral parameters like energy interaction parameters viz. Slater-Condon factor (Fk), Racah (Ek), Lande spin-orbit interaction (ζ4f), nephelauxetic ratio (β), bonding parameter (b1/2), per cent covalency (δ), and the intensity parameters like oscillator strength (P) and Judd-Ofelt Tλ, (λ ​= ​2,4,6) parameters, of Pr3+ ion complexes with reduced Glutathione (GSH) in the presence and absence of Mg2+ in different aqueous solutions of CH3OH, C4H8O2, CH3CN and DMF. The variations in the values of the energy interaction and intensity parameters clearly demonstrates the relative sensitivity of the 4f-4f transitions and its correlation with ligand structure and the nature of metal-ligand interaction. Further, the reaction dynamics and thermodynamic properties for the complexation of Pr3+ with glutathione and Mg2+ ligand have been investigated using different computed parameters like rate constant (k), activation energy (Ea), A (pre-exponential factor) and thermodynamic parameters, ΔH0, ΔG0 and ΔS0.  相似文献   

7.
From an analysis of the internal rate parameters of the pyrene excimer in solution in propylene glycol, isopropanol and other solvents it is concluded that (i) the radiative transition rate (kFD)0, normalized to a medium of n=1, is independent of the solvent, (ii) the temperature-independent radiationless transition rate kID0 is due to intersystem crossing, and (iii) the temperature-dependent radiationless transition rate kIDT is due to internal conversion, induced by thermally-activated molecular motion which is subject to viscous restraint by the solvent. For the crystal pyrene excimer the radiative transition rate = (kFD)0, kID0 = 0, and kIDT is due to intersystem crossing.  相似文献   

8.
The second osmotic virial coefficient (A2) and its entropic and enthalpic parts (A2,s and A2,H) have been determined, by means of light-scattering measurements, for solutions of polystyrene, polymethylmethacrylate and cellulose nitrate of different molecular weights in 19 solvents. A distinct qualitative correlation exists between A2 and A2,H and between A2,s and A2,H. The elimination of the “geometric” parameters of the polymer, by dividing these coefficients by suitably chosen reduction parameters, shows that the reduced coefficients obtained A20 and A2,s0 are predominantly functions of the reduced enthalpy coefficient A2,H0.  相似文献   

9.
The absorbance characteristics and influential factors on these characteristics for a liquid-phase gas sensor, which is based on gas–permeable liquid core waveguides (LCWs), are studied from theoretical and experimental viewpoints in this paper. According to theory, it is predicted that absorbance is proportional to the analyte concentration, sampling time, analyte diffusion coefficient, and geometric factor of this device when the depletion layer of the analyte is ignored. The experimental results are in agreement with the theoretical hypothesis. According to the experimental results, absorbance is time-dependent and increasing linearly over time after the requisite response time with a linear correlation coefficient r2 > 0.999. In the linear region, the rate of absorbance change (RAC) indicates improved linearity with sample concentration and a relative higher sensitivity than instantaneous absorbance does. By using a core liquid that is more affinitive to the analyte, reducing wall thickness and the inner diameter of the tubing, or increasing sample flow rate limitedly, the response time can be decreased and the sensitivity can be increased. However, increasing the LCW length can only enhance sensitivity and has no effect on response time. For liquid phase detection, there is a maximum flow rate, and the absorbance will decrease beyond the stated limit. Under experimental conditions, hexane as the LCW core solvent, a tubing wall thickness of 0.1 mm, a length of 10 cm, and a flow rate of 12 mL min−1, the detection results for the aqueous benzene sample demonstrate a response time of 4 min. Additionally, the standard curve for the RAC versus concentration is RAC = 0.0267 c + 0.0351 (AU min−1), with r2 = 0.9922 within concentrations of 0.5–3.0 mg L−1. The relative error for 0.5 mg L−1 benzene (n = 6) is 7.4 ± 3.7%, and the LOD is 0.04 mg L−1. This research can provide theoretical and practical guides for liquid–phase gas sensor design and development based on a gas-permeable Teflon AF 2400 LCW.  相似文献   

10.
In studies of solid supported lipid bilayers with atomic force microscopes (AFM) the force between tip and bilayer is routinely measured. During the approach of the AFM tip in aqueous electrolyte first a short-range repulsive force is observed. For many solid-like and some liquid-like lipid bilayers a subsequent break-through is observed. We observe such a break-through also for dioleoyloxypropyl-trimethylammonium chloride (DOTAP) which is expected to be liquid-like. Here we describe a model which assumes that the jump reflects the penetration of the AFM tip through the lipid bilayer. The model predicts a logarithmical dependence of the break-through force on the approaching velocity of the AFM tip. Two parameters are introduced: The ratio A/αV, α being a geometric factor, A being the area over which pressure is exerted on the bilayer, V the activation volume, and k0, the rate of spontaneous formation of a hole in the lipid bilayer that is big enough to allow the break-through of the tip. Experiments with bilayers consisting of DOTAP and dioleoylphosphatidylserine (DOPS) show that the break-through forces behave in the predicted way. For DOTAP we obtain ratios A/αV of about 58 nm−1 and rates k0 ranging from 1.9×10−8 to 2.5×10−4 s−1. For DOPS the corresponding values are 162 nm−1 and 2.0 s−1.  相似文献   

11.
The fluorescence lifetime of trans-stilbene in dilute methylcyclohexane/iso-hexane solution has been measured and the mean S1 radiative (kF), radiationless (kI) and cis-isomerization (kC) rate parameters have been determined from ?90 to 60°C. Si consists of a fluorescent trans (1Bu*) state (kF0 = 6.0 × 108 s?1) which undergoes reversible thermal-activated rotational internal conversion (ΔH = 1.75 kcal mole?1, ΔS = 10.6 cal deg?1 mole?1) to a non-fluorescent perp (1Ag*) state. p(1Ag*) lies 610 cm?1 above t (1Bu*) with an intermediate S1 potential maximum. p(1Ag*) undergoes internal conversion(kI. = 5.8 × 108 s?1) to p (1Ag) leading to cis-isomerization. This is the main isomerization channel over the whole temperature range.  相似文献   

12.
This work reports the study of the kinetics of zinc recovery from spent pickling solutions by means of emulsion pertraction technology (EPT) in order to reuse the metal in electrolytic processes. Tributyl phosphate (TBP) and service water were used as extraction (EX) and back-extraction (BEX) agents, respectively. Kinetic experiments were carried out in hollow fiber membrane contactors in order to analyse the influence of several operation variables on the rate of zinc recovery. A mathematical model that considers the mass transfer resistance shared between the organic liquid membrane and the organic phase boundary layer was developed; the mass transfer coefficients were estimated by means of the parameter estimation tool ASPEN CUSTOM MODELER (from ASPENTECH) to obtain the values km = 2.68 × 10−7 m/s and AVko = 0.0125 s−1. Simulated results agreed satisfactorily well with experimental data. Consequently, the kinetic model and parameters were confirmed. Finally, a comparison between EPT and non-dispersive solvent extraction (NDSX) was carried out in order to evaluate the advantages and disadvantages of both membrane configurations.  相似文献   

13.
The photoexcited triplet state of phenazine in toluene glasses at 35 K is investigated by light modulation-EPR spectroscopy. From the transient EPR spectra and the kinetics in the three canonical orientations (p = x, y, z) the rate parameters are determined. Thus, the depopulation rate constants kp, the anisotropic spin lattice relaxation rate constants Wp, and the ratios between the population constants Ap are calculated: kx = (2.2 ± 0.3) × 102 s?1, ky = (0.21 ± 0.04) × 102 s?1, kz = (0.06 ± 0.03) × 102 s?1, Wx = (8.6 ± 0.9) × 103 s?1, Wy = (11.0 ± 1.0) × 103 s?1, Wz = (14.0 ± 1.4) × 103 s?1, and Ax: Ay:Az ≈ 1:0.04:0.02. It is concluded therefore that the in-plane spin state |τx > is the active one.  相似文献   

14.
The kinetics of oxygen uptake in the cumyl peroxide-initiated oxidation of cyclohexanol (373 K, o-dichlorobenzene) is studied. The parameters of the oxidizability of k p (2k t )?0.5 (which depend on [RH]) and the rate constants of the bi- and trimolecular reactions of chain initiation (k 0 = 1.25 × 10?8 L/(mol s) and k0 = 2.5 × 10?9 L2/(mol2 s), respectively) are determined by solving the inverse kinetic problem. It is demonstrated that the quadratic-law recombination of peroxyl radicals during cyclohexanol oxidation also occurs without chain termination. The recombination rates of peroxyl radicals with and without chain termination (k′/k t ) are found to grow with increasing [RH], reaching their maxima at [RH] = 1.0 mol/L, and to diminish subsequently. We conclude that this can be attributed to changes in the ratio between the propagating peroxyl radicals (hydroperoxyl and 1-hydroxycyclohexylperoxyl) in the reaction medium.  相似文献   

15.
Rate constants kq for fluorescence quenching of eleven 9 and 9,10 substituted anthracences (Ai) by tris(pentafluorophenyl) phosphine (FP) are reported. Correlations of kq with the electronic properties of Ai and FP reveal that A+ FP charge-transfer interactions are important in the quenching process for several of the Ai. No one quenching mechanism is able to explain all of the data. The quenching mechanism of anthracene, 9-methyl-anthracene and 9,10-dimethylanthracene in benzene appears to lack a pathway available to anthracenes whose substituents contain a lone pair of electrons.  相似文献   

16.
Rate constants kq for fluorescence quenching of eleven 9 and 9,10 substituted anthracenes (Ai) by tris(pentafluorophenyl)phosphine (FP) are reported. Correlations of kq with the electronic properties of Ai and FP reveal that A+ FP charge-transfer interactions are important in the quenching process for several of the Ai. No one quenching mechanism is able to explain all of the data. The quenching mechanism of anthracene, 9-methyl-anthracene and 9,10-dimethylanthracene in benzene appears to lack a pathway available to anthracenes whose substituents contain a lone pair of electrons.  相似文献   

17.
The photodissociation of ketene, CH2CO(X?1A1) → CH21A1) + CO(X 1Σ+) has been observed at 337 nm, using a pulsed nitrogen laser. The CH21A1) radical has been detected by laser induced fluorescence with a tunable dye laser. A laser excitation spectrum has been obtained from CH21A1) over the wavelength interval from 588.9 to 595.6 nm in the Σ ← Π vibronic subband of the CH21A1); υ″ = 0, 0, 0?b? 1B1; υ′ = 0, 14, 0) transition. For the CH21A1 ; υ′= 0, 0, 0?X? 3B1; υ′' = 0, 0, 0) energy separation an upper limit of (6.3 ± 0.8) kcal/mole has been found. The radiative lifetime τ and the rate constant k for the removal of the 000 rotational level of the Σ(0, 14, 0) vibronic state have been measured directly. The values are τ = (4.2 ± 0.2) μs and k = (7.4 ± 0.3) × 10?10 cm3 molecule?1 s?1, respectively.  相似文献   

18.
In 2010 we investigated the applicability of the current k 0 and k 0-fission factors for the determination of the n(235U)/n(238U) isotopic ratio in multi-elemental samples containing uranium. An overestimation 3–4 % was observed in our determinations when employing the recommended 2003 k 0-literature. After a recalibration of all our laboratory instruments, a 3 % overestimation was still observed in this work when employing this nuclear data. Therefore we aimed at the experimental re-determination of these composite nuclear constants in order to enhance the reliability of the isotopic ratio determination method and the accuracy of our data-filtering algorithms. New k 0-fission factors are given for 7 nuclides that are not currently present in the 2012 k 0-database. Several additional k 0 factors are introduced for some nuclides in this library. Our k 0 results are also compared with those recently reported by Blaauw et al.  相似文献   

19.
The OPAL research reactor in Australia has been used to determine k 0 values for 134mCs, 134Cs, 192Ir and 194Ir. Values for 24Na have also been measured for quality control. The neutron flux at the irradiation positions was very highly thermalised (f > 2,000), resulting in almost negligible activation by epithermal neutrons. As a consequence, the contribution to the total uncertainty of the k 0 values from epithermal-related factors such as Q 0 and $ \bar{E}_{\text{r}} $ was very small. The measured caesium k 0 values have been compared with the library values as well as with recent measurements by St Pierre et al. and Farina Arboccò et al. While there are k 0 values for 194Ir in the library, no 192Ir values have been measured previously. Despite 192Ir having a higher sensitivity than 194Ir, k 0 values were not measured during the establishment of the k 0-method because the nuclear data available at the time indicated that the activation cross-section of 191Ir deviated significantly from 1/v behaviour (g(T n ) ≠ 1), which would result in unacceptable errors if k 0 analysis were to be carried out using the Høgdahl convention. However later nuclear data compilations showed that 191Ir has better 1/v behaviour than previously reported, making it suitable for k 0 analysis using the Høgdahl convention. For completeness, k 0 values have been determined using both the Høgdahl and modified-Westcott conventions and these have been compared with library (194Ir) and calculated values.  相似文献   

20.
The kinetics of oxidation of N,N′-ethylenebis(isonitrosoacetyleacetoneimine)copper(II) complex, CuIIL, by N-bromosuccinimide (SBr) in weakly aqueous acidic solutions was studied under pseudo-first-order conditions. Plots of ln(A  ? A t ) versus time where A t and A are absorbance values of the Cu(III) product at time t and infinity, respectively, showed marked deviations from linearity. The curves showed an acceleration of reaction rate consistent with an autocatalytic behavior. In the presence of Hg(II) ions, plots of ln(A  ? A t ) versus time are linear up to >85 % of reaction. The value of the observed rate constant, k obs, increases with decreasing pH. At constant reaction conditions, the dependence of the observed rate constants, k obs, is described by Eq. (1). 1 $$ k_{\text{obs}} = k_{\text{o}} + k_{1} \left[ {{\text{H}}^{ + } } \right] $$ The dependence of both k o and k 1 on [SBr] is not linear. The mechanism of the title reaction is consistent with an inner sphere mechanism in which a pre-equilibrium step precedes the electron transfer step. The overall rate law is represented by Eq. (2) where [CuIIL]t and K 1 represent the total copper(II) complex concentration and the pre-equilibrium formation constant, respectively. 2 $$ d\left[ {{\text{Cu}}^{\text{III}} {\text{L}}^{ + } } \right]/dt = \left\{ {\left( {k_{\text{o}} + k_{1} \left[ {{\text{H}}^{ + } } \right]} \right)\left[ {\text{SBr}} \right]\left[ {{\text{Cu}}^{\text{II}} {\text{L}}} \right]_{t} } \right\}/\left( {1 + K_{1} \left[ {\text{SBr}} \right]} \right) $$ .  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