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.
The 57Fe Mössbauer effect in two samples (A and B) of [Fe(papt)2] and in its solvates with CHCl3 and C6H6 has been studied between 4.2 and 343 K and clearly indicates a temperature induced high-spin (5T2) ? low-spin (1A1) transition in these compounds [paptH = 2-(2-pyridylamino)-4-(2-pyridyl) thiazole]. At 343 K, sample B shows a doublet with ΔEQ = 2.03 mm s?1 and δIS = +0.87 mm s?1, characteristic of a 5T2 ground state. At 257 K, a second doublet, typical for a 1A1 ground state, is observed and its intensity increases as the transition progresses but levels off below ~ 100 K. At 4.2 K, 83% of the intensity is due to the 1A1 state, and ΔEQ(1A1) = 1.56 mm s?1 and δIS(1A1 = +0.32 mm s?1. In an applied magnetic field, Vzz(1A1) < 0 and η ≈ 0.7 have been determined, whereas for the sT2 ground state, Vzz(sT2) > 0, η ≈ 0.75, and an internal hyperfine field Hn ≈ ?13 kG have been observed. Similar results have been obtained with the other samples.Debye-Waller factors f5T2 and f1A1 were determined from the saturation corrected areas in the Mössbauer spectra, assuming Curie-Weiss dependence of the magnetic susceptibility for the 5T2 and constant υcff for the 1A1 ground state. The temperature dependence of ?In f1A1 closely follows the Debye model with Θ1A1 = 165 K, whereas the same applies to ?ln f5T2 only above ~ 210 K and Θ5T2 = 134 K. The nature of the observed transition is discussed and the data presented are shown to be incompatible with a model based on a Boltzmann distribution between the two states.  相似文献   

2.
Spectral simulation was used to analyze the molecular rovibrational bands of D2H and H2D at 5600 Å. These bands were previously measured by the ion beam neutralization method. They were assigned to the electronic 3p2B1 ? 2s2A1 and vibrational (ν - ν″) = (0, 0, 0,-0, 0, 0) transitions. Least squares fits to the experimental line-positions were made to determine the asymmetric rotator constants A, B and C for the 2s2A1 and 3p2B1 ν = 0 states of D2H and H2D, hitherto unknown. Lorentz line-profiles were assumed for the D2H and H2D rotational lines, whose widths are mainly governed by the lifetimes of the lower states. The bands at 5600 Å were simulated and the 2s2A1 state lifetimes were estimated to be σ ≥ 0.5 ± 0.2 ps for D2H and σ ≥ 0.4 ± 0.2 ps for H2D. Vibrational constants of D3 and D2H in the 2s2A1 states are determined from the positions of the 0-0 and 0-1 vibrational bands given in respective experimental spectra previously measured. For the first time the vibrational constants ω1 and ω2 of the 2s2A1 state of H2D were estimated from the positions of the 0-0 and 0-1 band maxima. These vibrational constants are compared with the corresponding vibrational constants of their ions.  相似文献   

3.
Donnan dialysis with an ion exchange membrane was investigated for ions of different valence. The effective diffusion coefficients (De) of various kinds of ions in the membrane were obtained by fitting of the equation derived from the Nernst–Planck equation to three or more sets of experimental data for Donnan dialysis. It became apparent that the value of De/Ds of monovalent ions (e.g., K+ or Na+ ions) at zA=1 and zB=2 (feed ions are monovalent ones and driving ions are bivalent ones) remained constant at ca. 1/210 and that of bivalent ions (e.g., Ca2+, Cu2+, or Mg2+ ions) remained constant at ca. 1/526 where Ds denotes the diffusion coefficient of ions at infinite dilution in water calculated from the Nernst–Einstein equation, and zA and zB represent the valences of the feed and driving ions, respectively. De/Ds of monovalent ions (e.g., H+, K+, or Na+ ions) at zA=2 and zB=1 (feed ions are bivalent ones and driving ions are monovalent ones) was constant at ca. 1/23.3 and that of bivalent ions remained constant at ca. 1/58.4. It was proved that De/D using De at zA=1 and zB=2 was constant at 1/3.0 and that at zA=2 and zB=1 remained constant at 3.0 where D represents the diffusion coefficient of ions in the membrane at zA=zB (the valences of both feed and driving ions are equal). Therefore, it was found that a large flux of ions could be obtained using the monovalent driving ions in Donnan dialysis. On the other hand, the small flux can be obtained using bi- or higher-valent driving ions.  相似文献   

4.
The dissociative ionization of 4,4-dimethyl-1-thia-4-silacyclohexane (I) and 2,3,3-trimethyl-1-thia-3-silacyclopentane(II) has been studied by electron photoionization (PI) mass spectrometric methods. The molecular ion fragmentation is mainly related to the loss of ethylene and results in a [Me2SiSC2H4]+? (m/z 118) ion-radical (A). Further loss of ethylene from A produces a dimethylsilanethione [Me2SiS]+? (m/z 90) ion-radical (B). The latter is the most abundant ion in the mass spectra of I and II at 70 eV.The ionization energies (IE) of I (8.22 ± 0.07 eV) and II (8.06 ± 0.03 eV) and the appearance energies (AE) of ion-radicals A and B have been determined. Also, the following heats of formation were calculated (kJ/mol): ΔHf0(I) = ?31.1; ΔHf0(II) = ?65.8; ΔHf0(MI+?) = 762.0; ΔHf0(MII+?)= 712.1; ΔHf0(A)aver = 780.2; ΔHf0(B)aver = 847.7.  相似文献   

5.
A novel complex [Ni(H2O)4(TO)2](NO3)2·2H2O (TO = 1,2,4-triazole-5-one) was synthesized and structurally characterized by X-ray crystal diffraction analysis. The decomposition reaction kinetic of the complex was studied using TG-DTG. A multiple heating rate method was utilized to determine the apparent activation energy (E a) and pre-exponential constant (A) of the former two decomposition stages, and the values are 109.2 kJ mol?1, 1013.80 s?1; 108.0 kJ mol?1, 1023.23 s?1, respectively. The critical temperature of thermal explosion, the entropy of activation (ΔS ), enthalpy of activation (ΔH ) and the free energy of activation (ΔG ) of the initial two decomposition stages of the complex were also calculated. The standard enthalpy of formation of the new complex was determined as being ?1464.55 ± 1.70 kJ mol?1 by a rotating-bomb calorimeter.  相似文献   

6.
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   

7.
In this paper a new relation between the second virial coefficients A_2, (?)_w and (dV_(es)/dC)_c→0=K_s was derived from proposed model theory of concentration effects in GPC for mono-and poly-dispersed polymers. Based on this relation a new method for determination of second vifial coefficients from the combination of (dV_(es)/dC)_c→0=K_3, (?)_w and K_H measurements was proposed.The values of A_2 for mono-and poly-dispersed polystyrenes with molecular weight range from 10~4 to 10~6 in good and theta solvents were determined by proposed method. Results show that their values of A_2 are in agreement with those obtained by light scattering.  相似文献   

8.
The use of the complex acid HAlCl4 (HCl+AlCl3) permits the detemrination of the standard potential of the hydrogen electrode in nitromethane. The result (E0(Hs+/H2)=0.5 V vs. Fc/Fc+, Fc=ferrocene) shows that nitromethane is very weakly basic. This measurement is confirmed by showing that the standard potential of the hydrogen electrode in various solvents is linked to Gutmann's donor numbers of these solvents. The E0(Hs+/H2) value obtained in nitromethane belongs to the correlation line.  相似文献   

9.
The aluminium(III)–citrate complex (NH4)4[Al2(C6H4O7)(C6H5O7)2]·4H2O was characterized using anion exchange chromatography on-line coupled with the element specific ICP-AES detector. Time-dependent monitoring of individual species in aqueous solution at different temperatures gave information about the species stability and the decomposition pathway. The aluminium–citrate complex (NH4)4[Al2(C6H4O7)(C6H5O7)2]·4H2O disintegrated via an unknown intermediary Al(III)–citrate species from which the thermodynamically stable complex [Al3(C6H4O7)3(OH)(H2O)]4− was formed. The activation energy for the decomposition reaction and the pre-exponential factor were determinated to be Ea = 81.95 kJ mol−1 and A = 3.62 × 1013 s−1.  相似文献   

10.
The mean activity coefficients for CsCl in N-methylformamide or urea (w) + H2O (1 ? w) systems were determined in this work by potentiometry, using ion-selective electrodes at 298.15 K. The value of mass fraction w was varied between 0.00 and 0.40 in five unit-steps and the molality of CsCl was between 0.0020 and 1.4009 mol·kg?1. The experimental data have been correlated with the Pitzer, modified Pitzer and the extended Debye–Hückel equations. The resulting values of the mean activity coefficients, the osmotic coefficients and the standard Gibbs energy of transfer, together with the Pitzer ion-interaction parameters (β (0), β (1) and C φ), extended Debye–Hückel parameters (a, c and d), and modified Pitzer parameters (b, B MX, C MX) are reported for the investigated systems.  相似文献   

11.
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   

12.
For measuring solution enthalpies of the strong hygroscopic double chlorides, an isoperibol solution calorimeter was built. Samples of 2–5 g could be solved to a molar dilution 1 : 3000. The temperature difference between reaction vessel and thermostat was measured by a thermopile; the temperature of the thermostat was constant to 2 · 10?4°C. From the molar enthalpies of solution (ΔHL), enthalpies for the reactions nACl + MCl2 = AsMCl(s + 2) were calculated: ΔHPR = ? ΔHL (double chloride) + ΔHL(n Cl) + ΔHL(MCL2) These values are relatively small: about ? 50 kJ for the Cs-compounds, nearly zero for the K- and Na-compounds.  相似文献   

13.
Molar conductance of lithium acetate, sodium acetate and potassium acetate were studied in aqueous 2-butanol solutions with an alcohol mass fraction (w2) of 0.70, 0.80 and 0.90 at 298.15, 303.15 and 308.15 K. The conductance data were analyzed with the Fuoss conductance-concentration equation to evaluate the limiting molar conductances (Λ0), association constants (KA,c) and cosphere diameter (R) for ion-pair formation. Gibbs energy (ΔG0), enthalpy (ΔH0) and entropy (ΔS0) for ion-association reaction were derived from the temperature dependence of KA,c. Activation energy for ionic movement (ΔH#) was derived from the temperature dependence of Λ0. Based on the composition dependence of Walden products (Λ0η0) and different thermodynamic properties (ΔG0H0, ΔS0 and ΔH#), the influence of the solvent composition on ion-association and solvation behavior of ions were discussed in terms of ion-solvent, ion-ion interactions and the structural changes in the mixed solvent media.  相似文献   

14.
Low-temperature heat capacities of a solid complex Zn(Val)SO4·H2O(s) were measured by a precision automated adiabatic calorimeter over the temperature range between 78 and 373 K. The initial dehydration temperature of the coordination compound was determined to be, T D=327.05 K, by analysis of the heat-capacity curve. The experimental values of molar heat capacities were fitted to a polynomial equation of heat capacities (C p,m) with the reduced temperatures (x), [x=f (T)], by least square method. The polynomial fitted values of the molar heat capacities and fundamental thermodynamic functions of the complex relative to the standard reference temperature 298.15 K were given with the interval of 5 K. Enthalpies of dissolution of the [ZnSO4·7H2O(s)+Val(s)] (Δsol H m,l 0) and the Zn(Val)SO4·H2O(s) (Δsol H m,2 0) in 100.00 mL of 2 mol dm–3 HCl(aq) at T=298.15 K were determined to be, Δsol H m,l 0=(94.588±0.025) kJ mol–1 and Δsol H m,2 0=–(46.118±0.055) kJ mol–1, by means of a homemade isoperibol solution–reaction calorimeter. The standard molar enthalpy of formation of the compound was determined as: Δf H m 0 (Zn(Val)SO4·H2O(s), 298.15 K)=–(1850.97±1.92) kJ mol–1, from the enthalpies of dissolution and other auxiliary thermodynamic data through a Hess thermochemical cycle. Furthermore, the reliability of the Hess thermochemical cycle was verified by comparing UV/Vis spectra and the refractive indexes of solution A (from dissolution of the [ZnSO4·7H2O(s)+Val(s)] mixture in 2 mol dm–3 hydrochloric acid) and solution A’ (from dissolution of the complex Zn(Val)SO4·H2O(s) in 2 mol dm–3 hydrochloric acid).  相似文献   

15.
Hydroboration reactions of 1-octene and 1-hexyne with H2BBr·SMe2 in CH2Cl2 were studied as a function of concentration and temperature, using 11B NMR spectroscopy. The reactions exhibited saturation kinetics. The rate of dissociation of dimethyl sulfide from boron at 25 °C was found to be (7.36 ± 0.59 and 7.32 ± 0.90) × 10−3 s−1 for 1-octene and 1-hexyne, respectively. The second order rate constants, k2, for hydroboration worked out to be 7.00 ± 0.81 M s−1 and 7.03 ± 0.70 M s−1, while the overall composite second order rate constants, k K, were (3.30 ± 0.43 and 3.10 ± 0.37) × 10−2 M s−1, respectively at 25 °C. The entropy and enthalpy values were found to be large and positive for k1, whilst for k2 these were large and negative, with small values for enthalpies. This is indicative of a limiting dissociative (D) for the dissociation of Me2S and associative mechanism (A) for the hydroboration process. The overall activation parameters, ΔH and ΔS, were found to be 98 ± 2 kJ mol−1 and +56 ± 7 J K−1 mol−1 for 1-octene whilst, in the case of 1-hexyne these were found out to be 117 ± 7 kJ mol−1 and +119 ± 24 J K−1 mol−1, respectively. When comparing the kinetic data between H2BBr·SMe2 and HBBr2·SMe2, the results showed that the rate of dissociation of Me2S from H2BBr·SMe2 is on average 34 times faster than it is in the case of HBBr2·SMe2. Similarly, the rate of hydroboration with H2BBr·SMe2 was found to be on average 11 times faster than it is with HBBr2·SMe2. It is also clear that by replacing a hydrogen substituent with a bromine atom in the case of H2BBr·SMe2 the mechanism for the overall process changes from limiting dissociative (D) to interchange associative (Ia).  相似文献   

16.
Experimental differential cross sections for 40 keV electrons scattered by C2H2, C2H4 and C2H6 molecules were measured using the gas electron diffraction method in the range of the scattering variable s from s = 1 A?1 to s = 30 A?1. The differential cross sections for neon were also measured and compared with calculated differential cross sections to calibrate the diffractograph. Experimental differential cross sections show significant deviations with respect to theoretical differential cross sections calculated from the Debye-Ehrenfest model, mainly in the range of small scattering angles. The observed differences are connected to chemical binding effects. From the experimental data, an estimation of the binding energy was carried out. The deduced values: ?0.58 ± 0.20 au for C2H2, ?0.94 ± 0.30 au for C2H4 and ?1.23 ± 0.40 au for C2H6 are in agreement with those obtained by thermochemical methods.  相似文献   

17.
The exothermic H-atom abstraction reaction of SO+2 with H2 has been studied in a selected ion flow drift tube (SIFDT) over a range of center-of-mass energies from thermal (300 K) to about 0.12 eV. The measured rate coefficient at 300 K is 4.2 × 10−12 cm3 s−1 which is very much less than the Langevin capture rate. The increase in rate coefficient with ion kinetic energy gives a linear Arrhenius-type plot with a slope that indicates a barrier of ∼5 kJ mol−1 exists on the potential surface. The H2SO+2 potential surface is also explored in an ab initio investigation using the G2 procedure. An (SO+2.H2)1 transition state between reactants and products is identified, corresponding to the barrier found from experiments.  相似文献   

18.
The diffusive gradients in thin films (DGT) technique, utilizing resin gel with ion-exchange resin Duolite GT73 and new ion-exchange resin Ambersep GT74, was investigated for the accumulation of four mercury species (Hg2+, CH3Hg+, C2H5Hg+, C6H5Hg+). The diffusion coefficients of mercury species in agarose gel calculated on the basis of Fick’s Law were mercury species-specific. The diffusion coefficients of Hg2+ and CH3Hg+ at 25 °C (9.07 ± 0.23 × 10−6 cm2 s−1 and 9.06 ± 0.30 × 10−6 cm2 s−1, respectively) were very similar, but the diffusion coefficients of C2H5Hg+ (6.87 ± 0.23 × 10−6 cm2 s−1) and C6H5Hg+ (3.86 ± 0.19 × 10−6 cm2 s−1) were significantly lower. Influence of experimental conditions (pH, selected cations, chlorides and humic substance) on mercury species accumulation by DGT was studied. The DGT technique was applied to river water spiked with mercury species.  相似文献   

19.
《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   

20.
We report two macrocyclic ligands based on a 1,7-diaza-12-crown-4 platform functionalized with acetate (tO2DO2A2−) or piperidineacetamide (tO2DO2AMPip) pendant arms and a detailed characterization of the corresponding Mn(II) complexes. The X−ray structure of [Mn(tO2DO2A)(H2O)]·2H2O shows that the metal ion is coordinated by six donor atoms of the macrocyclic ligand and one water molecule, to result in seven-coordination. The Cu(II) analogue presents a distorted octahedral coordination environment. The protonation constants of the ligands and the stability constants of the complexes formed with Mn(II) and other biologically relevant metal ions (Mg(II), Ca(II), Cu(II) and Zn(II)) were determined using potentiometric titrations (I = 0.15 M NaCl, T = 25 °C). The conditional stabilities of Mn(II) complexes at pH 7.4 are comparable to those reported for the cyclen-based tDO2A2− ligand. The dissociation of the Mn(II) chelates were investigated by evaluating the rate constants of metal exchange reactions with Cu(II) under acidic conditions (I = 0.15 M NaCl, T = 25 °C). Dissociation of the [Mn(tO2DO2A)(H2O)] complex occurs through both proton− and metal−assisted pathways, while the [Mn(tO2DO2AMPip)(H2O)] analogue dissociates through spontaneous and proton-assisted mechanisms. The Mn(II) complex of tO2DO2A2− is remarkably inert with respect to its dissociation, while the amide analogue is significantly more labile. The presence of a water molecule coordinated to Mn(II) imparts relatively high relaxivities to the complexes. The parameters determining this key property were investigated using 17O NMR (Nuclear Magnetic Resonance) transverse relaxation rates and 1H nuclear magnetic relaxation dispersion (NMRD) profiles.  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