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 234 毫秒
1.
A standard enthalphy of -767.2 kJ mol-1 was determined for sulphuryl fluoride from the heat of alkaline hydrolysis, and a standard free energy of -720.8 kJ mol-1 derived. The thermodynamically preferred product of ionic fluorosulphate decompositions is always sulphuryl fluoride. Kinetic control must predominate when sulphur trioxide is the main product.  相似文献   

2.
Oxygen equilibria of haemoglobin were analysed according to a binding isotherm proposed by Amire ( Bull. Chem. Soc. Jpn. 1994, 67, 7 )1 to obtain the intrinsic oxygen association constants to the molecule. Two sets of binding sites in haemoglobin were identified, which were ascribed to R2 and T forms of the molecule. The average intrinsic association constants determined as a function of temperature gave a heat of oxygenation of‐76 ± 4 kJ mol?;1 (tetramer). A microcalorimetrically determined heat of deoxygenation of oxyhaemoglobin by dithionite gave ?267 ± 10 kJ mol?1 (tetramer). From these results, the heat of allostery of ?234 ± 24 kJ mol?1 for haemoglobin tetramer was obtained, yielding allosteric energy per salt bridge of‐29 ± 3 kJ. This result suggests that salt‐bridge may, in fact, be thermochemically equivalent to hydrogen bonds.  相似文献   

3.
In an effort to understand the reaction of antibiotic hydrolysis with B2 metallo-??-lactamases (M??Ls), the thermodynamic parameters of imipenem hydrolysis catalyzed by metallo-??-lactamase ImiS from Aeromonas veronii bv. sobria were determined by microcalorimetric method. The values of activation free energy $ \Updelta G_{ \ne }^{\theta } $ are 86.400?±?0.043, 87.543?±?0.034, 88.772?±?0.024, and 89.845?±?0.035?kJ?mol?1 at 293.15, 298.15, 303.15, and 308.15?K, respectively, activation enthalpy $ \Updelta H_{ \ne }^{\theta } $ is 18.586?±?0.009?kJ?mol?1, activation entropy $ \Updelta S_{ \ne }^{\theta } $ is ?231.34?±?0.12?J?mol?1?K?1, apparent activation energy E is 21.084?kJ?mol?1, and the reaction order is 1.5. The thermodynamic parameters reveal that the imipenem hydrolysis catalyzed by metallo-??-lactammase ImiS is an exothermic and spontaneous reaction.  相似文献   

4.
Kinetics of polyurethane formation between several polyols and isocyanates with dibutyltin dilaurate (DBTDL) as the curing catalyst, were studied in the bulk state by differential scanning calorimetry (DSC) using an improved method of interpretation. The molar enthalpy of urethane formation from secondary hydroxyl groups and aliphatic isocyanates is 72±3 kJ mol-1 and for aromatic isocyanates it is 55±2 kJ mol-1 . In the case of a single second order reaction for aliphatic isocyanates reaction, activation energy is 70±5 kJ mol-1 with oxypropylated polyols and 50±3 kJ mol-1 with Castor oil. For aromatic isocyanates and oxypropylated polyols the activation energy is higher around 77 kJ mol-1 . In the case of two parallel reactions (situation for IPDI and TDI 2-4) best fits are observed considering two different activation energies.  相似文献   

5.
The surface segregation of In and S from a dilute Cu(In,S) ternary alloy were measured using Auger electron spectroscopy coupled with a linear programmed heater. The alloy was linearly heated and cooled at constant rates. Segregation data of a linear heat run showed surface segregation of In that reached a maximum surface coverage of 25% followed by S, which reached a coverage of 30%. It was found that after In had reached a maximum surface coverage, it started to desegregate as soon as the S enriched the surface until In was completely replaced by S. The segregation parameters, namely, the pre‐exponential factor (D0), activation energy (Q), segregation energy (ΔG?) and interaction energy (Ω) were extracted from the measured segregation data for both In and S segregation in Cu by simulating the measured segregation data with a theoretical segregation model (modified Darken model). The segregation parameters obtained for In segregation in Cu are D0 = 1.8 ± 0.5 × 10?5 m2 s?1, Q = 184.3 ± 1.0 kJ.mol?1, ΔG? = ?61.4 ± 1.4 kJ.mol‐1, ΩCu?In = 3.0 ± 0.4 kJ.mol?1; for S segregation in Cu the parameters are D0 = 8.9 ± 0.5 × 10?3 m2 s?1, Q = 212.8 ± 3.0 kJ.mol?1, ΔG? = ?120.0 ± 3.5 kJ.mol?1, ΩCu?S = 23.0 ± 2.0 kJ mol?1 and the In and S interaction parameter is ΩIn?S = ?4.0 ± 0.5 kJ.mol?1. The initial parameters used for the Darken calculations were extracted from fits performed with the Fick's and Guttmann model. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

6.
The ionization energies and [C3H5O]+ appearance energies for a series of oxygenated organic compounds have been measured by dissociative photoionization mass spectrometry. The adiabatic ionization energy for cyclopentanol is observed to be 9.72 eV. A 298 K heat of formation of 591.2±2.3kJ mol?1, based on the stationary electron convention, is derived for the propanoyl cation in the gas phase. A heat of formation of –86±6 kJ mol?1 is obtained for methylketene, which leads to an absolute proton affinity of 853±8 kJ mol?1.  相似文献   

7.
Knudsen effusion studies of the sublimation of polycrystalline SnSe and SnSe2, prepared by annealing and chemical vapor transport reactions, respectively, have been carried out using vacuum microbalance techniques in the temperature ranges 736–967 K and 608–760 K, respectively. From experimental mass-loss data for the sublimation reaction SnSe(s) = SnSe(g), the recommended values for the heat of formation and absolute entropy of SnSe(s) were calculated to be ΔH°298,f = ?86.4 ± 9.9 kJ · mol?1 and S°298 = 89.0 ± 7.1 J · K?1 · mol?1. From mass-loss data for the decomposition reaction \documentclass{article}\pagestyle{empty}\begin{document}$ {\rm SnSe}_{\rm 2} ({\rm s)} = {\rm SnSe(s)} + \frac{1}{{\rm x}}{\rm Se}_{\rm x} ({\rm g) (x} = 2 - 8) $\end{document}, the recommended values for the heat of formation and absolute entropy of SnSe2(s) were determined to be ΔH°298,f = ?118.1 ± 15.1 kJ · mol?1 and S°298 = 111.8 ± 11.8 J · K?1 mol?1.  相似文献   

8.
The kinetics of the interactions between three sulfur‐containing ligands, thioglycolic acid, 2‐thiouracil, glutathione, and the title complex, have been studied spectrophotometrically in aqueous medium as a function of the concentrations of the ligands, temperature, and pH at constant ionic strength. The reactions follow a two‐step process in which the first step is ligand‐dependent and the second step is ligand‐independent chelation. Rate constants (k1 ~10?3 s?1 and k2 ~10?5 s?1) and activation parameters (for thioglycolic acid: ΔH1 = 22.4 ± 3.0 kJ mol?1, ΔS1 = ?220 ± 11 J K?1 mol?1, ΔH2 = 38.5 ± 1.3 kJ mol?1, ΔS2 = ?204 ± 4 J K?1 mol?1; for 2‐thiouracil: ΔH1 = 42.2 ± 2.0 kJ mol?1, ΔS1 = ?169 ± 6 J K?1 mol?1, ΔH2 = 66.1 ± 0.5 kJ mol?1, ΔS2 = ?124 ± 2 J K?1 mol?1; for glutathione: ΔH1 = 47.2 ± 1.7 kJ mol?1, ΔS1 = ?155 ± 5 J K?1mol?1, ΔH2 = 73.5 ± 1.1 kJ mol?1, ΔS2 = ?105 ± 3 J K?1 mol?1) were calculated. Based on the kinetic and activation parameters, an associative interchange mechanism is proposed for the interaction processes. The products of the reactions have been characterized from IR and ESI mass spectroscopic analysis. A rate law involving the outer sphere association complex formation has been established as   相似文献   

9.
From measurement of the heat of hydrolysis, at 25°C , the enthalpy of formation of rubidium tetrafluoroiodate is derived: ΔH°f [RbIF4, cryst.]298= ?191.12±4.43 kJ mol?1. Heat capacity measurements for RbIF4 over the range 273–303 K are also reported.  相似文献   

10.
The rate of the fastest ene reaction between 4-phenyl-1,2,4-triazoline-3,5-dione (1) and 2,3-dimethyl-2-butene (2) is studied by means of stopped flow in solutions of benzene (k 2 = 55.6 ± 0.5 and 90.5 ± 1.3 L mol?1 s?1 at 23.3 and 40°C) and 1,2-dichloroethane (335 ± 9 L mol?1 s?1 at 23.5°C). The enthalpy of reaction (?139.2 ± 0.6 kJ/mol in toluene and ?150.2 ± 1.4 kJ/mol in 1,2-dichloroethane) and the enthalpy (20.0 ± 0.5 kJ/mol) and entropy (144 ± 2 J mol?1 K?1) of activation are determined. A clear correlation is observed between the reaction rate and ionization potential in a series of ene reactions of 4-phenyl-1,2,4-tri-azoline-3,5-dione with acyclic alkenes.  相似文献   

11.
To accurately derive the kinetic and thermodynamic parameters governing the hydrolysis of the lactone ring at physiological pH, a derivative spectrophotometric technique was used for the simultaneous estimation of lactone and carboxylate forms of camptothecin (CPT). The hydrolysis of the CPT‐lactone and the lactonization of CPT‐carboxylate at 310.15 K followed a first‐order decay with apparent rate constants equal to 0.0279 ± 0.0016 min?1 and 0.0282 ± 0.0024 min?1, respectively. The activation energy associated with the hydrolysis of the CPT‐lactone and the lactonization of the CPT‐carboxylate as calculated from the Arrhenius equation was 89.18 ± 0.84 and 86.49 ± 2.7 kJ mol?1, respectively. The enthalpy and entropy of the thermodynamically favored hydrolysis reaction were on average 10.49 kJ mol?1 and 48.00 J K?1 mol?1, respectively. The positive enthalpy and entropy values of the CPT‐lactone hydrolysis indicate that the reaction is endothermic and entropically driven. The stability of CPT‐lactone in the presence of human serum albumin (HSA) was also analyzed. Notwithstanding the much faster hydrolysis of the CPT‐lactone in the presence of HSA at various temperatures, the energy of activation was determined to be similar to the one estimated in the absence of HSA, suggesting that HSA does not catalyze the hydrolysis reaction, but it merely binds, sequesters, and stabilizes the CPT‐carboxylate species. © 2009 Wiley Periodicals, Inc. Int J Chem Kinet 41: 704–715, 2009  相似文献   

12.
At room temperature and below, the proton NMR spectrum of N-(trideuteriomethyl)-2-cyanoaziridine consists of two superimposed ABC patterns assignable to two N-invertomers; a single time-averaged ABC pattern is observed at 158.9°C. The static parameters extracted from the spectra in the temperature range from –40.3 to 23.2°C and from the high-temperature spectrum permit the calculation of the thermodynamic quantities ΔH0 = ?475±20 cal mol?1 (?1.987 ± 0.084 kJ mol?1) and ΔS0 = 0.43±0.08 cal mol?1 K?1 (1.80±0.33 J mol?1 K?1) for the cis ? trans equilibrium. Bandshape analysis of the spectra broadened by non-mutual three-spin exchange in the temperature range from 39.4–137.8°C yields the activation parameters ΔHtc = 17.52±0.18 kcal mol?1 (73.30±0.75 kJ mol?1), ΔStc = ?2.08±0.50 cal mol?1 K?1 (?8.70±2.09 J mol?1 K?1) and ΔGtc (300 K) = 18.14±0.03 kcal mol?1 (75.90±0.13 kJ mol?1) for the transcis isomerization. An attempt is made to rationalize the observed entropy data in terms of the principles of statistical thermodynamics.  相似文献   

13.
The complex formation reactions of [Cu(NTP)(OH2)]4? (NTP?=?nitrilo-tris(methyl phosphonic acid)) with some selected bio-relevant ligands containing different functional groups, are investigated. Stoichiometry and stability constants for the complexes formed are reported. The results show that the ternary complexes are formed in a stepwise mechanism whereby NTP binds to copper(II), followed by coordination of amino acid, peptide or DNA. Copper(II) is found to form Cu(NTP)H n species with n?=?0, 1, 2 or 3. The concentration distribution of the various complex species has been evaluated. The kinetics of base hydrolysis of glycine methyl ester in the presence of copper(II)-NTP complex is studied in aqueous solution at different temperatures. It is proposed that the catalysis of GlyOMe ester occurs by attack of OH? ion on the uncoordinated carbonyl carbon atom of the ester group. Activation parameters for the base hydrolysis of the complex [Cu(NTP)NH2CH2CO2Me]4? are, ΔH±?=?9.5?±?0.3?kJ?mol?1 and ΔS±?=??179.3?±?0.9?J?K?1?mol?1. These show that catalysis is due to a substantial lowering of ΔH±.  相似文献   

14.
The ionization energies and [C4H9]+ appearance energies for several 2-substituted butanes have been measured by photoionization mass spectrometry. Using the stationary electron convention, a 298 K heat of formation of 771±3 kJ mol?1 is derived for the sec-butyl cation in the gas phase. A value of 747±3 kJ mol?1 is calculated for the proton affinity of trans-2-butene.  相似文献   

15.
The kinetics of the reactions of ground state oxygen atoms with 1-pentene, 1-hexene, cis-2-pentene, and trans-2-pentene was investigated in the temperature range 200 to 370 K. In this range the temperature dependences of the rate constants can be represented by k = A′ Tn exp(− E′a/RT) with A′ = (1.0 ± 0.6) · 10−14 cm3 s−1, n = 1.13 ± 0.02, E′a = 0.54 ± 0.05 kJ mol−1 for 1-pentene: A′ = (1.3 ± 1.2) · 10−14 cm3 s−1, n = 1.04 ± 0.08, E′a = 0.2 ± 0.4 kJ mol−1 for 1-hexene; A′ = (0.6 ± 0.6) · 10−14 cm3 s−1, n = 1.12 ± 0.05, E′a = − 3.8 ± 0.8 kJ mol−1 for cis-2-pentene; and A′ = (0.6 ± 0.8) · 10−14 cm3 s−1, n = 1.14 ± 0.06, E′a = − 4.3 ± 0.5 kJ mol−1 for trans-2-pentene. The atoms were generated by the H2-laser photolysis of NO and detected by time resolved chemiluminescence in the presence of NO. The concentrations of the O(3P) atoms were kept so low that secondary reactions with products are unimportant. © 1997 John Wiley & Sons, Inc.  相似文献   

16.
Tetraethoxysilane (TEOS) is widely used to synthesize siliceous material by the sol–gel process. However, there is still some disagreement about the nature of the limiting step in the hydrolysis and condensation reactions. The goal of this research was to measure the variation in the concentration of intermediates formed in the acid-catalyzed hydrolysis by 29Si NMR spectroscopy, to model the reactions, and to obtain the rate constants and the activation energy for the hydrolysis and early condensation steps. We studied the kinetics of TEOS between pH 3.8 and 4.4, and four temperature values in the range of 277.2–313.2?K, with a TEOS:ethanol:water molar ratio of 1:30:20. Both hydrolysis and the condensation rate speeded up with the temperature and the concentration of oxonium ions. The kinetic constants for hydrolysis reactions increased in each step kh1?<?kh2?<?kh3?<?kh4, but the condensation rate was lower for dimer formation than for the formation of the fully hydrolyzed Si(OH)4. The system was described according to 13 parameters: six of them for the kinetic constants estimated at 298.2?K, six to the activation energies, and one to the equilibrium constant for the fourth hydrolysis. The mathematical model shows a steady increase in the activation energy from 34.5?kJ?mol?1 for the first hydrolysis to 39.2?kJ?mol?1 in the last step. The activation energy for the condensation reaction from Si(OH)4 was ca. 10?kJ?mol?1 higher than the largest activation energy in the hydrolytic reactions. The decrease in the net positive charge on the Si atom contributes to the protonation of the ethoxy group and makes it a better leaving group.  相似文献   

17.
Restricted rotation about the naphthalenylcarbonyl bonds in the title compounds resulted in mixtures of cis and trans rotamers, the equilibrium and the rotational barriers depending on the substituents. For 2,7-dimethyl-1,8-di-(p-toluoyl)-naphthalene (1) ΔH° = 3.66 ± 0.14 kJ mol?1, ΔS° = 1.67 ± 0.63 J mol?1 K?1, ΔHct = 55.5 ± 1.3 kJ mol?1, ΔHct = 51.9 ± 1.3 kJ mol?1, ΔSct = ?41.3±4.1 J mol?1 K?1 and ΔSct = ?42.9±4.1 J mol?1 K?1. The rotation about the phenylcarbonyl bond requires ΔH = ?56.9±4.4 kJ mol?1 and ΔS = ?20.5±15.3 J mol?1 K?1 for the cis rotamer, and ΔH = 43.5Δ0.4 kJ mol?1 and ΔS =± ?22.4Δ1.3 J mol?1 K?1 for the trans rotamer. The role of electronic factors is likely to be virtually the same for both these rotamers but steric interaction between the two phenyl rings occurs in the cis rotamer only. Hence, the difference of the activation enthalpies obtained for the cis and trans rotamers, ΔΔH?1 = 13.4 kJ mol?1, provides a basis for the estimation of the role of steric factors in this rotation. For the tetracarboxylic acid 2 and its tetramethyl ester 3 the equilibrium is even more shifted towards the trans form because of enhanced steric and electrostatic interactions between the substituents in the cis form. The barriers for the rotation around the phenylcarbonyl bond and the cis-trans isomerization are lowered; an explanation for this result is presented.  相似文献   

18.
The rates of the electron self‐exchange between uranyl(VI) and uranyl(V) complexes in solution have been investigated in detail with quantum chemical methods. The calculations have shown that the bond length of U? Oyl is elongated by 0.1 Å when the extra electron is localized on the sites. The diabatic potential surfaces are obtained. The inner reorganization energies are 212.6 and 226.8 kJ mol?1 for hydroxide and fluoride bridge systems, respectively. The solvent reorganization energies are 28.12 and 31.60 kJ mol?1 for hydroxide and fluoride bridge systems, respectively. The nuclear frequency factors are 3.17 × 1013 and 3.12 × 1013 s?1 for hydroxide and fluoride bridge systems, respectively. The electronic coupling matrix elements are 1.89 and 4.06 kJ mol?1 for hydroxide and fluoride bridge systems, respectively. The electron‐transfer rates of our calculations are 12.95 and 0.819 M?1 s?1 for hydroxide and fluoride bridge systems, respectively. © 2009 Wiley Periodicals, Inc. Int J Quantum Chem, 2010  相似文献   

19.
The gas‐phase elimination kinetics of the above‐mentioned compounds were determined in a static reaction system over the temperature range of 369–450.3°C and pressure range of 29–103.5 Torr. The reactions are homogeneous, unimolecular, and obey a first‐order rate law. The rate coefficients are given by the following Arrhenius expressions: ethyl 3‐(piperidin‐1‐yl) propionate, log k1(s?1) = (12.79 ± 0.16) ? (199.7 ± 2.0) kJ mol?1 (2.303 RT)?1; ethyl 1‐methylpiperidine‐3‐carboxylate, log k1(s?1) = (13.07 ± 0.12)–(212.8 ± 1.6) kJ mol?1 (2.303 RT)?1; ethyl piperidine‐3‐carboxylate, log k1(s?1) = (13.12 ± 0.13) ? (210.4 ± 1.7) kJ mol?1 (2.303 RT)?1; and 3‐piperidine carboxylic acid, log k1(s?1) = (14.24 ± 0.17) ? (234.4 ± 2.2) kJ mol?1 (2.303 RT)?1. The first step of decomposition of these esters is the formation of the corresponding carboxylic acids and ethylene through a concerted six‐membered cyclic transition state type of mechanism. The intermediate β‐amino acids decarboxylate as the α‐amino acids but in terms of a semipolar six‐membered cyclic transition state mechanism. © 2005 Wiley Periodicals, Inc. Int J Chem Kinet 38: 106–114, 2006  相似文献   

20.
From hydrolysis and solution measurements the enthalpies of formation of SbF5(?), LiSbF6(s), NaSbF6(s), KSbF6(s), CsSbF6(s), AgSbF6(s), and SbF6?aq. are estimated to be ?1324 ± 12, ?2062 ± 5, ?2060 ± 6, ?2080 ± 3, ?2082 ± 15, ?1653 ± 3, and ?1789 ± 4 kJ mol?1 respectively. Less precise estimates of the enthalpies of formation of O2SbF6 and of CsSb3F16 are also given. From the results the fluoride ion affinity of SbF5, the single ion hydration enthalpy of SbF6? (g), and the charge distribution within the SbF6? ion have been calculated.  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