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 718 毫秒
1.
Mixtures of 2-ethylhexylsodium and 2-ethylhexyllithium are studied by 1H- and 13C-NMR spectroscopy in the temperature range from 20 to −50°C in hydrocarbon solutions. Characteristic temperature-dependent spectra obtained are indicative of dynamic exchange processes taking place in the system. The following activation parameters are found: ΔH=31.7±2.7 kJ mol−1; ΔG313=58.7±0.6 kJ mol−1; ΔS=−86.37±10.8 J mol−1 K−1. The negative value of the activation entropy indicates that the exchange proceeds through the associative mechanism. The participation in exchange reactions of aggregates, containing both sodium and lithium derivatives, is suggested.  相似文献   

2.
The rate of decomposition of H2O2 in the presence of Fe(III)-y complex (y is ethylenebis(oxyethylenedinitrilo)tetraacetic acid (EGTA) anion) was investigated under variable conditions of pH and temperature, various water-miscible solvents, and different concentrations of H2O2, [Fe-y], and acetate ions. The following rate law holds: Rate = (k1K3K4/[H+]) [Fe-y(OH)]2− [H2O2] at pH less than 9.80, and Rate = (k2K5[H+]/K3) [Fe-y(OH)2]3−[OOH] at pH above 9.80. The values of k1K4and k2K5 at 25 °C were found to be 1523 and 0.747 M−1 S−1, respectively. Activation enthalpy and activation entropy for this reaction were determined from Arrhenius plots and found to be ΔH* = 34.38 K J mol−1 and ΔS* = −167.2 J K−1 mol−1.  相似文献   

3.
The kinetics of the reaction of the CH3CHBr, CHBr2 or CDBr2 radicals, R, with HBr have been investigated in a temperature-controlled tubular reactor coupled to a photoionization mass spectrometer. The CH3CHBr (or CHBr2 or CDBr2) radical was produced homogeneously in the reactor by a pulsed 248 nm exciplex laser photolysis of CH3CHBr2 (or CHBr3 or CDBr3). The decay of R was monitored as a function of HBr concentration under pseudo-first-order conditions to determine the rate constants as a function of temperature. The reactions were studied separately from 253 to 344 K (CH3CHBr + HBr) and from 288 to 477 K (CHBr2 + HBr) and in these temperature ranges the rate constants determined were fitted to an Arrhenius expression (error limits stated are 1σ + Student’s t values, units in cm3 molecule−1 s−1, no error limits for the third reaction): k(CH3CHBr + HBr) = (1.7 ± 1.2) × 10−13 exp[+ (5.1 ± 1.9) kJ mol−1/RT], k(CHBr2 + HBr) = (2.5 ± 1.2) × 10−13 exp[−(4.04 ± 1.14) kJ mol−1/RT] and k(CDBr2 + HBr) = 1.6 × 10−13 exp(−2.1 kJ mol−1/RT). The energy barriers of the reverse reactions were taken from the literature. The enthalpy of formation values of the CH3CHBr and CHBr2 radicals and an experimental entropy value at 298 K for the CH3CHBr radical were obtained using a second-law method. The result for the entropy value for the CH3CHBr radical is 305 ± 9 J K−1 mol−1. The results for the enthalpy of formation values at 298 K are (in kJ mol−1): 133.4 ± 3.4 (CH3CHBr) and 199.1 ± 2.7 (CHBr2), and for α-C–H bond dissociation energies of analogous compounds are (in kJ mol−1): 415.0 ± 2.7 (CH3CH2Br) and 412.6 ± 2.7 (CH2Br2), respectively.  相似文献   

4.
The kinetics of the interaction of thiosemicarbazide with cis-[Ru(bipy)2(H2O)2]2+ (bipy = α α′-bipyridyl) have been studied spectrophotometrically as a function of [Ru(bipy)2(H2O)22+], [bipyridyl] and temperature, at a particular pH (4.8), where the substrate complex exists predominantly as the diaqua species and thiosemicarbazide as the neutral ligand. The reaction proceeds via an outer sphere association complex formation, followed by two slow consecutive steps. The first is the conversion of the aforementioned complex into the inner sphere complex, and the second step involves the entrance of another thiosemicarbazide molecule in the coordination zone of Ru(II) whereby, in each step, an aqua ligand is replaced. The association equilibrium constant (KE) for the outer sphere complex formation has been evaluated together with rate constants for the two subsequent steps. Activation parameters have been calculated for both steps using the Eyring equation (ΔH1# = 25.37±1.6 kJ mol−1, ΔS1# = −215.48 ± 4.5 J K−1 mol−1, ΔH2# = 24.24 ± 1.1 kJ mol−1, ΔS2# = −207.14 ± 3.0 J K−1 mol−1). The low enthalpy of activation and large negative value of entropy of activation indicate an associative mode of activation for both aqua ligand substitution processes. From the temperature dependence of KE, the thermodynamic parameters calculated are: ΔH0 = 10.75±0.54 kJ mol−1 and ΔS0 = 84.67 ± 1.75 J K−1 mol−1, which give a negative ΔG0 value at all temperatures studied, supporting the spontaneous formation of an outersphere association complex prior to the first step.  相似文献   

5.
Kinetics of Formation of Peroxyacetic Acid   总被引:1,自引:0,他引:1  
The kinetics of the reaction of acetic acid with hydrogen peroxide, leading to peroxyacetic acid, were studied at various molar reactant ratios (AcOH-H2O2 from 6 : 1 to 1 : 6) at 20, 40, and 60°C and sulfuric acid (catalyst) concentrations of 0 to 9 wt %. The reaction is reversible, and the equilibrium constant decreases as the temperature rises: K = 2.10 (20°C), 1.46 (40°C), 1.07 (60°C); Δr H 0 = − 13.7±0.1 kJ mol−1, Δr S = −40.5±0.4 J mol−1 K−1. The maximal equilibrium concentration of peroxyacetic acid (2.3 M) is attained at 20°C and a molar AcOH-to-H2O2 ratio of 2.5 : 1. The rate constants of both forward and reverse reactions increase with increase in sulfuric acid concentration from 0 to 5 wt %. Further raising the catalyst concentration does not affect the reaction rate. The reaction mechanism is discussed.__________Translated from Zhurnal Obshchei Khimii, Vol. 75, No. 7, 2005, pp. 1187–1193.Original Russian Text Copyright © 2005 by Dul’neva, Moskvin.  相似文献   

6.
Summary A ternary solid complex Gd(Et2dtc)3(phen) has been obtained from reactions of sodium diethyldithiocarbamate (NaEt2dtc), 1,10-phenanthroline (phen) and hydrated gadolinium chloride in absolute ethanol. The title complex was described by chemical and elemental analyses, TG-DTG and IR spectrum. The enthalpy change of liquid-phase reaction of formation of the complex, ΔrHΘm(l), was determined as (-11.628±0.0204) kJ mol-1 at 298.15 K by a RD-496 III heat conduction microcalorimeter. The enthalpy change of the solid-phase reaction of formation of the complex, ΔrHΘm(s), was calculated as (145.306±0.519) kJ mol-1 on the basis of a designed thermochemical cycle. The thermodynamics of reaction of formation of the complex was investigated by changing the temperature of liquid-phase reaction. Fundamental parameters, the apparent reaction rate constant (k), the apparent activation energy (E), the pre-exponential constant (A), the reaction order (n), the activation enthalpy (ΔrHΘ), the activation entropy (ΔrSΘ), the activation free energy (ΔrGΘ) and the enthalpy (ΔrHΘ), were obtained by combination of the thermodynamic and kinetic equations for the reaction with the data of thermokinetic experiments. The constant-volume combustion energy of the complex, ΔcU, was determined as (-18673.71±8.15) kJ mol-1 by a RBC-II rotating-bomb calorimeter at 298.15 K. Its standard enthalpy of combustion, ΔcHΘm, and standard enthalpy of formation, ΔfHΘm, were calculated to be (-18692.92±8.15) kJ mol-1 and (-51.28±9.17) kJ mol-1, respectively.  相似文献   

7.
We have studied 18 reactions, including four identity reactions, involving transfer of a dimethylcarbamoyl group with N-acylpyridinium bonds to pyridine and its 4-substituted derivatives in acetonitrile solutions at 298 K. The rate constants k ij varied within the range 0.4 to 10–6 L/mol·s; the equilibrium constants K ij varied from 107 to 10–5. The rate and equilibrium for exchange of carbamoyl groups are described satisfactorily by the Brönsted equation. We have shown that all the reactions occur according to a forced concerted S N2 mechanism. The structure of the transition state and its position on the reaction coordinate for identity transfer are considered using a More O'Ferrall-Jencks diagram.  相似文献   

8.
The relative mobilities of the nitro group and fluorine atom in 1,3-dinitrobenzene and 1-fluoro-3-nitrobenzene by the action of phenols in the presence of potassium carbonate in dimethylformamide at 95–125°C were studied by the competing reaction method. The rate constant ratios k(NO2)/k(F) were correlated with the differences between the corresponding activation parameters (ΔΔH and ΔΔ S ). The greater mobility of the nitro group was found to be determined by the entropy control of the reactivity of arenes. The activation parameters (ΔH and ΔS ) were calculated, and the enthalpy-entropy compensation effect was revealed. The reaction mechanism is discussed.__________Translated from Zhurnal Organicheskoi Khimii, Vol. 41, No. 7, 2005, pp. 999–1005.Original Russian Text Copyright © 2005 by Khalfina, Vlasov.  相似文献   

9.
10.
The solubilization of pyrene in aqueous solution of β-cyclodextrin (β-CD) or its derivatives such as β-CD-hexanoyl, β-CD-benzoyl and β-CD-dodecylsulfonate was investigated by spectrophotometry. Linear and non-linear regression methods were used to estimate the association constants (K1). A 1:1 stoichiometric ratio and different effects of the hexanoyl, benzoyl and dodecylsulfonate groups on the association constant were observed for the binary inclusion complex between pyrene and β-CD. The formation constant was shown to decrease when β-CD was modified by a dodecylsulfonate chain. The value of K1 was 190 ± 10 L mol−1 for the [pyrene/β-CD] complex and 145 L mol−1 for the [pyrene/β-CD-dodecylsulfonate] complex. Partitioning of the pyrene molecules between the dodecylsulfonate chains and cyclodextrin cavities can explain the decrease in the association constant value. In the cases of β-CD-hexanoyl and β-CD-benzoyl derivatives, no association constants were detected. Results suggest that the high hydrophobicity of the hexanoyl and benzoyl groups prevents the inclusion of pyrene molecules inside the cyclodextrin cavity.  相似文献   

11.
The pressure of thermal dissociation of platinum tetrachloride by the first step PtCl4(s) = PtCl3(s) + 0.5 Cl2(g) was measured by the static method with a quartz membrane-gauge zero-pressure manometer. An approximating equation for the dissociation pressure vs. temperature was found. The enthalpy (52160±880 J mol−1) and entropy (72.1±1.6 J mol−1 K−1) of dissociation were calculated. The heat of formation found for platinum tetrachloride (−246.3±1.3 kJ mol−1) at 298.15 K agrees well with the value obtained by the calorimetric method (−245.6±1.9 kJ mol−1).__________Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 10, pp. 2028–2031, October, 2004.  相似文献   

12.
Enthalpies for the two proton ionizations of the biochemical buffers N-tris(hydroxymethyl)methyl-4-aminobutanesulfonic acid (TABS), N-tris(hydroxymethyl)methyl-3-aminopropanesulfonic acid (TAPS) and 3-[N-tris(hydroxymethyl)methylamino]-2-hyroxypropane sulfonic acid (TAPSO) were obtained in water–methanol mixtures with methanol mole fraction (Xm) from 0 to 0.360. The ionization enthalpy for the first proton (ΔH1) of all three buffers was small and exhibited slight changes upon methanol addition. The ionization enthalpy of the second proton (ΔH2) of TABS increased from 39.6 to 49.8 kJ mol−1 and for TAPS from 40.1 to 43.2 kJ mol−1, with a minimum of 38.2 kJ mol−1 at Xm = 0.059. For TAPSO the increase was from 33.1 to 35.6 kJ mol−1 at Xm = 0.194, with measurements at higher Xm precluded by low solubility of TAPSO in methanol rich solvents. The solvent composition was selected so as to include the region of maximum structure enhancement of water by methanol. The results were interpreted in terms of solvent–solvent and solvent–solute interactions.  相似文献   

13.
The electronic structure and stability of pyrrolyl are investigated using CASSCF, CASPT2 and G2(MP2) techniques. The ground state of pyrrolyl is found to be 2A2, with five π-electrons, as in cyclopentadienyl. The computed N–H bond energy of pyrrole is 94.8 kcal mol−1, while the heat of formation ΔfH298o of pyrrolyl is deduced to be 70.5±1 kcal mol−1. The Arrhenius parameters of N–H and C–H bond fission in pyrrole and cyclopentadiene and hydrogen abstraction reactions (by hydrogen) were also computed, indicating that pyrrolyl forms predominantly by C–H bond fission of pyrrolenine rather than by direct N–H bond fission.  相似文献   

14.
The oxidation of pyruvic acid by the title silver(III) complex in aqueous acidic (pH, 1.1–4.5) media is described. The reaction products are MeCO2H and CO2, together with a colourless solution of the Ag+ ion. The free ligand, ethylenebis(biguanide) is released in near-quantitative yield upon completion of the reduction. The parent complex, [Ag(H2L)]3+ and one of its conjugate bases, [Ag(HL)]2+, participate in the reaction with both pyruvic acid (HPy) and the pyruvate anion (Py) as the reactive reducing species. Ag+ was found to be catalytically inactive. At 25.0°C, I=1.0moldm–3, rate constants for the reactions [Ag(H2L)]3++HPy (k 1), [Ag(H2L)]3++Py (k 2), [Ag(HL)]2++HPy (k 3) and [Ag(HL)]2++Py (k 4) arek 1=(94±6)×10–5dm3mol–1s–1, (k 2 K a+k 3 K a1)= (1.3±0.1)×10–5s–1 and k 4=(58±4)×10–5dm3mol–1s–1, respectively, where K a1is the first acid dissociation constant of the [Ag(H2L)]3+ and K a is for pyruvic acid. A comparison between the k 1 and k 4 values is indicative of the judgement that k 2k 3. A one-electron inner-sphere redox mechanism seems more justified than an outer-sphere electron-transfer between the redox partners.  相似文献   

15.
The singlet—triplet energy splitting (ΔE ST = E S E T ) was calculated for formylnitrene (5) and for the syn- and anti-rotamers of carboxynitrene HOC(O)N (6) by the CCSD(T) method. Extrapolation of ΔE ST to a complete basis set was calculated to be negative for 5 and strongly positive for 6. Similar results were obtained by the G2 procedure. The reason for the dramatic stabilization of the singlet state appeared to be a special bonding interaction between the nitrogen and oxygen atoms, which results in the structure intermediate between those of nitrene and oxazirene. It was found that the B3LYP/6-31G(d) method overestimates ΔE ST by ∼9 kcal mol−1 for 5 and by ∼7 kcal mol−1 for 6. Taking into account this overestimation and the results of DFT calculations, it was concluded that benzoylnitrene has a singlet ground state. It was proved experimentally using photolysis of benzoyl azide in an argon matrix at 12 K that benzoylnitrene has a singlet ground state and its structure is similar to that of oxazirene. Nevertheless, these singlet intermediates have low barrier to the aziridine formation, which is traditionally considered to be a typical singlet nitrene reaction.__________Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 3, pp. 519–526, March, 2005.  相似文献   

16.
BAFP (2,6-bis[4-(4-amino-2-trifluoromethylphenoxy)benzoyl] pyridine), a synthesized polyimide compound, was exploited for the first time to analyze its interaction with human serum albumin (HSA) by molecular modeling, fluorescence and Fourier transform infrared attenuated total reflection spectroscopy (FTIR ATR) with drug concentrations of 3.3 × 10−6 to 3.0 × 10−5 mol L−1. Molecular docking was performed to reveal the possible binding mode. The results suggested that BAFP can strongly bind to human serum albumin (HSA) and the primary binding site of BAFP is located in site II of HSA, which is supported by the results from the competitive experiment. The binding constants for the interaction of BAFP with HSA have been evaluated from relevant fluorescence data at different temperatures (296, 303, 310 and 308 K). The alterations of the protein secondary structure in the presence of BAFP in aqueous solution were quantitatively calculated by the evidences from FTIR ATR spectroscopes. The binding process was exothermic and spontaneous, as indicated by the thermodynamic analyses, and the major part of the binding energy is hydrophobic interaction, which is also in good agreement with the results of molecule modeling study. The enthalpy change ΔH0, the free energy change ΔG0 and the entropy change ΔS0 of 296 K were calculated to be −7.75, −27.68 kJ mol−1 and 67.33 J mol−1 K−1, respectively.  相似文献   

17.
The interaction of molecular hydrogen with [Rh(PPh3)3]+ (1a) “immobilized” in the interlamellar spaces of montmorillonite resulted in the formation of a monohydrido complex, [RhIIH(PPh3)3] (2a), characterized by electrochemical data of the clay-loaded electrode, IR, EPR and hydrogen absorption studies. Heterogenized homogeneous catalytic hydrogenation of cyclohexene catalysed by 1a was investigated in the temperature range 283–313 K. The order of reaction with respect to cyclohexene and hydrogen concentration is fractional and first order with respect to catalyst concentration. Thermodynamic parameters ΔH0 and ΔS0 corresponding to the formation of the monohydrido species were found to be 18 kcal mol−1 and 61 e.u., respectively. The activation enthalpy, ΔH, and entropy, ΔS, for the hydrogenation of cyclohexene by the RhII—H complex in clay are more negative by about 2 kcal mol−1 and 7 e.u. compared to Wilkinson's catalyst, RhCl(PPh3)3 (1), in homogeneous solution.  相似文献   

18.
The reagent 2-pyridyl thiourea produces a reddish brown complex with rhenium in hydrochloric acid medium in the presence of tin(II) chloride. The complex shows absorption maxima at 405 nm, obeys Beer's law for a range of 1–16 ppm of rhenium. The sensitivity and molar absorptivity is found to be 0.0115g cm–2 and 1.603×104 1 mol–1 cm–1 respectively showing an improvement over other sulphur-nitrogen bearing ligands. Composition found by Job's and mole ratio method, indicates that the complex contains metal and reagent in the ratio 12. Stability constant and stepwise formation constants of the complex have been evaluated by Harvey-Manning's method (logK overall=8.825), Leden's method (logK overall=8.3096), Rossotti-Rossotti's method (logK overall=8.653) and Yatsimirskii's method (logK overall=8.740). The relative error per 1% absolute photometric error is found to be 2.7%.  相似文献   

19.
Ion-pair formation constants (mol–1 dm3 unit), KMX for a univalent metal salt (MX) and KMLX for its ion-pair complex (ML+X) with a crown ether (L) in water, were determined at various ionic strengths (I) and 25°C by potentiometry with ion-selective electrodes for MX=NaPic, NaMnO4, NaBPh4, KPic, and KMnO4; and MLX=Na(18C6)Pic, K(18C6)Pic, and Na(18C6)BPh4, where Pic and 18C6 denote a picrate ion and 18-crown-6 ether, respectively. Equations for analyzing I-dependence of logKMLX and logKMX were derived and fitted well to the I-dependence using a non-linear regression analysis. The equilibrium constants at I=0 mol dm–3, KMLX° and KMX°, were simultaneously obtained from the analysis. The experimental values of KMLX and KMX were only in agreement with the values calculated from KMLX° and KMX°, respectively, in the ranges of higher I.  相似文献   

20.
The kinetics of the silver(I)-catalysed autoxidation of SO3 2– into SO4 2– in ammonia–ammonium nitrate buffer obeyed the rate law:R obs=k1 k2 K[AgI]T[SO3 2-}][O2] / ([NH3]+K[SO3 2-])(k1+k2[O2])The values of k 1, k 2/k –1 and K were found to be 1.2l mol–1 s–1, 5.3 × 102 l mol–1 and 0.6 respectively at 30 °C. Two alternative free radical mechanisms have been proposed.  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