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.
The kinetics of acid-catalyzed hydrolysis of the [Co(en)(L)2(O2CO)]+ ion (L = imidazole, 1-methylimidazole, 2-methylimidazole) follows the rate law –d[complex]/dt = {k 1 K[H+]/(1 + K[H+])}[complex] (15–30 or 25–40 °C, [H+] = 0.1–1.0 M and I = 1.0 M (NaClO4)). The reaction course consists of a rapid pre-equilibrium protonation, followed by a rate determining chelate ring opening process and subsequent fast release of the one-end bound carbonato ligand. Kinetic parameters, k 1 and K, at 25 °C are 5.5 × 10–2 s–1, 0.44 M–1 (ImH), 5.1 × 10–2 s–1, 0.54 M–1 (1-Meim) and 3.8 × 10–3 s–1, 0.74 M–1 (2-MeimH) respectively, and activation parameters for k 1 are H1 = 43.7 ± 8.9 kJ mol–1, S1 = –123 ± 30 J mol–1 deg–1 (ImH), H1 = 43.1 ± 0.3 kJ mol–1, S1 = –125 ± 1 J mol–1 deg–1 (1-Meim) and H1 = 64.2 ± 4.3 kJ mol–1, S1 = –77 ± 14 J mol–1 deg–1 (2-MeimH). The results are compared with those for similar cobalt(III) complexes.  相似文献   

2.
Summary The vibrational spectra of solutions have been analyzed to assess both qualitatively and quantitatively the changes in enthalpy and entropy for ion pair formation in solutions of LiNCS, Mg(NCS)2, and LiN3 in liquid ammonia, dimethylformamide, dimethylsulphoxide and acetonitrile. Contrary to predictions both the H ass and S ass terms are all positive in the cases examined, indicating that the driving force in the ion association process derives from solvent-solute restructuring, and not the energy of the interaction between the cation and anion. This characteristic of contact ion pair formation is likely to be found to be applicable over a wide range of solvents. The following specific values of the thermodynamic parameters at 298 K have been obtained: LiNCS/DMF, G=–1.3 (1) kJ mol–1, H ass =+1.8 (5) kJ mol, S ass =+10 (2) J mol–1 K–1; LiNCS/DMSO, G=+0.9 (2) kJ mol–1, H ass =+0.3 (3) kJ mol–1; Mg(NCS)2/DMF, G ass =–4.0 (3) kJ mol–1, H ass =+15 (4) kJ mol–1, S=+64 (17) kJ mol–1; LiN3/DMSO, G ass =–2.5 (3) kJ mol–1, H ass =+4.9 (9) kJ mol–1, S ass =+25 (10) J K–1 mol–1.Submitted to celebrate the 70th Birthday of Professor Viktor Gutmann, and in recognition of his considerable contributions towards the better understanding of Chemistry in the Solution Phase  相似文献   

3.
Activation parameters of the interconversion of geometric isomers6a and6b were determined by a complete lineshape analysis of the temperature-dependent13C NMR spectra of 7,8-dipropyl-7-borabicyclo[4.2.2]deca-2,4,9-triene (6). For the reaction6a 6b, G 298 = 52.2±0.1 kJ mol–1, H = 27.9±0.5 kJ mol–1, S = –82±8 J mol–1 K–1; For the reaction6b 6a, G 298 = 52.6±0.1 kJ mol–1, H = 24.7±0.5 kJ mol–1, S = –93±10 J mol–1 K–1. The interconversion of deuteropyridine complexes9a and9b proceedsvia their dissociation, which indicates that the rearrangement of borane6 occurs according to the [1,3]-B shift mechanism.Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 9, pp. 2243–2250, September, 1996.  相似文献   

4.
Temperature dependence was studied for relative quantum yields of emission from some exciplexes of pyrene, 1,12-benzoperylene, and 9-cyanoanthracene with methoxybenzenes or methylnaphthalenes in solvents of different polarity (ranging from toluene to acetonitrile). The enthalpy H Ex *, the entropy S Ex *, and the Gibbs free energy G Ex *of formation of the exciplexes were determined. Depending of the Gibbs free energy of excited-state electron transfer (G et *) and solvent polarity, the values of H Ex *, S Ex *, and G Ex *vary over the ranges from –5 to –40 kJ mol–1, from +3 to –90 J mol–1K–1, and from +3 to –21 kJ mol–1, respectively. The possibility is discussed that the effect of solvent polarity G et *on the exciplex formation enthalpies can be rationalized in terms of the model of correlated polarization of an exciplex and the medium.  相似文献   

5.
Summary Concerning the relation between the experimental heat of fusion H* and the specific volumev of PETP a considerable uncertainty exists in literature. For PBTP obviously no data have been reported. The present paper reports H* andv measurements for undrawn PETP and PBTP samples which have been crystallized from the glassy state or from the melt at different temperatures for different periods of time.For PETP a linear relation is obtained: H* = 1411–1886v (Jg–1). Published values for the specific volumev c of the PETP crystal range from 0.660 to 0.687 cm3g–1. Ifv c = 0.660 cm3g–1 is accepted, a heat of fusion M m = 166 Jg–1 is obtained for the PETP crystal.For PBTP also a linear relation is found: H* = 1296–1628v (Jg–1). Withv c = 0.71 cm3g–1 one obtains H M = 140 Jg–1 as the heat of fusion of the PBTP crystal. The specific volumev a of amorphous PBTP (H* = 0) is 0.796 cm3g–1 which is much higher than the hitherto used values of 0.781–0.782 cm3g–1. The reason for this difference is thatv a cannot directly be measured, because the low quasi-static glass temperature of 15 °C enables quenched PBTP to undergo cold crystallization at 20 °C.
Zusammenfassung Hinsichtlich des Zusammenhangs zwischen experimenteller Schmelzwärme H* und spezifischem Volumenv von PETP bestehen in der Literatur beträchtliche Diskrepanzen. Für PBTP wurden bislang offensichtlich keine Ergebnisse veröffentlicht. In der vorliegenden Arbeit werden Messungen von H* undv für unverstreckte PETP- und PBTP-Proben mitgeteilt, die unterschiedlich lange bei ver-schiedenen Temperaturen aus dem Glaszustand oder aus der Schmelze kristallisiert wurden.Für PETP ergibt sich die lineare Beziehung: H* = 1411–1886v (Jg–1). Literaturwerte für das spezifische Volumenv c des PETP-Kristalls schwanken zwischen 0.660 und 0.687 cm3g–1. Nimmt manv c = 0.660 cm3g–1 als richtig an, so erhält man als Schmelzwärme des PETP-Kristalls H M = 166 Jg–1 = 32 kJ mole–1.Auch für PBTP erhält man eine lineare Abhängigkeit: H* = 1296–1628v. Mitv c = 0.71 cm3g–1 ergibt sich als Schmelzwärme des PBTP-Kristalls H M = 140 Jg–1 = 31 kJ mole–1. Das spezifische Volumen des amorphen PBTP beträgt a = 0.796 cm3g–1 und ist erheblich größer als der bisher angenommene Wert von 0.781 cm3g–1. Die Ursache fÜr diese Diskrepanz liegt darin begündet, daßv a nicht direkt gemessen werden kann, weil wegen der niedrigen quasi-statischen Glastemperatur von 15°C bei abgeschrecktem PBTP die Kaltkristallisation bei 20°C bereits einsetzt.


With 7 figures and 3 tables

Dedicated to Professor Dr. Matthias Seefelder on the occasion of his 60th birthday  相似文献   

6.
Summary G2 theory is shown to be reliable for calculating isodesmic and homodesmotic stabilization energies (ISE and HSE, respectively) of benzene. G2 calculations give HSE and ISE values of 92.5 and 269.1 kJ mol–1 (298 K), respectively. These agree well with the experimental HSE and ISE values of 90.5±7.2 and 268.7±6.3 kJ mol–1, respectively. We conclude that basis set superposition error corrections to the enthalpies of the homodesmotic or isodesmic reactions are not necessary in calculations of the stabilization energies of benzene using G2 theory. The calculated values of the enthalpies of formation of such molecules containing multiple bonds such as benzene ands-trans 1,3-butadiene, which are found from the enthalpies of isodesmic and homodesmotic reactions rather than of atomization reactions, demonstrate good performance of G2 theory. Estimates of theH f o value for benzene from the G2 calculated enthalpies of homodesmotic reaction (2) and isodesmic reaction (3) are 80.9 and 82.5 kJ mol–1 (298 K), respectively. These are very close to the experimentalH f o value of 82.9±0.3 kJ mol–1. TheH f o value ofs-trans 1,3-butadiene calculated using the G2 enthalpy of isodesmic reaction (4) is 110.5 kJ mol–1 and is in excellent agreement with the experimentalH f o value of 110.0±1.1 kJ mol–1.  相似文献   

7.
The title reaction has been studied spectrophotometrically in aqueous medium as a function of [substrate complex], [ligand], pH and temperature at constant ionic strength. At the physiological pH (7.4) the interaction with azide shows two distinct consecutive steps, i.e., it shows a non-linear dependence on the concentration of N3 ; both processes are [ligand]-dependent. The rate constant for the processes are: k 110–3 s–1 and k 210–5 s–1. The activation parameters calculated from Eyring plots are: H 1 = 14.8 ± 1 kJ mol–1, S 1 = –240 ± 3 J K–1 mol–1, H 2 = 44.0 ± 1.5 kJ mol–1 and S 2 = –190 ± 4 J K–1 mol–1. Based on the kinetic and activation parameters an associative interchange mechanism is proposed for the interaction process. From the temperature dependence of the outersphere association equilibrium constant, the thermodynamic parameters calculated are: H 1 0 = 4.4 ± 0.9 kJ mol–1, S 1 0 = 64 ± 3 J K–1 mol–1 and H 2 0 = 14.2 ± 2.9 kJ mol–1, S 2 0 = 90 ± 9 J K–1 mol–1, which gives a negative G 0 value at all temperatures studied, supporting the spontaneous formation of an outersphere association complex.  相似文献   

8.
The kinetics of base hydrolysis of the trans-[Cr(NH3)2(NCS)4] anion follows the rate law: -d[complex]/dt = k 0 + k 1[OH] (50–70 °C, [OH] = 0.1–1.9 M and = 2.0 M). The specific salt effect has been investigated for eight aqueous media: NaCl, NaBr, NaI, NaClO4, KCl, KBr, CsCl and CsBr. The alkali-independent path (k 0) does not show any specific effect of inert electrolyte ions, the activation parameters: H = 113.5 ± 0.4 kJ mol–1 and S = 24.1 ± 1.3 J mol–1 K–1 are interpreted in the frame of a dissociative interchange mechanism (I d). For the alkali-dependent path (k 1) the specific salt effect is observed for cations of the inert electrolyte, showing an important role for ion-pair formation between the cations and reagent complex anion in the activation process. A linear correlation between lnk 1 and lnK 0 (K 0 – ion-pair formation constant) has been found for the cations studied. The dissociative, via conjugate base, mechanism (D CB) has been proposed for the alkali-dependent path.  相似文献   

9.
The sublimation pressure of chromium trichloride was measured by the static method with a quartz membrane-gauge manometer in the temperature range of 875–1230 K. An approximating equation for the sublimation pressure vs. temperature was found. The enthalpy (259.4±4 kJ mol–1) and the entropy (224.2±3.5 J mol–1 K–1) of sublimation at 298 K were calculated. For the process 2 CrCl3(g) + Cl2(g) = 2 CrCl4(g), the following values were obtained: r H°298 = –207.1±11.6 kJ mol–1 and r S°298 = –173.6±10 5 J mol–1 K–1.Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 8, pp. 1561–1564, August, 2004.  相似文献   

10.
The thermochemical study of cubane-1,4-dicarboxylic acid (1), diethyl cubane-1,4-dicarboxylate (2), diisopropyl cubane-1,4-dicarboxylate (3), and bis(2-fluoro-2,2-dinitro)ethyl cubane-1,4-dicarboxylate (4) was performed. The standard enthalpies of combustion (c H°) and formation (f H°) of these compounds were estimated using the method of combustion in a calorimetric bomb in an oxygen atmosphere. Using the additive group method, calculated values for f H° of these substances which agreed satisfactorily with the experimental ones were obtained. The strain energies (E s) of the cubic structure of derivatives1–4 were calculated. It was concluded thatE s did not change on substitution of hydrogen atoms in cubane for various functional groups and was equal toE s of the structure of cubane itself. The reliability of the single published value of f H° in the cubane crystal state, 541.8 kJ mol–1 (129.5 kcal mol–1), was confirmed.Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 10, pp. 2471–2473, October, 1996.  相似文献   

11.
Summary Kinetic studies of the anation of the title complex by NO 2 show that it occurs in a stepwise manner leading to thecis-dinitro-complex both steps having a common rate equation:-d[complex]/dt = a[NO 2 ]/{[NO 2 ] + b}. The variation ofpseudo-first-order rate constant (kobs) with [NO 2 ] indicates that the reaction proceeds through ion-pair interchange path. Activation parameters calculated by the Eyring equation are: H 1 = (65±7) kJ mol–1 and S 1 = (–82±11) JK–1 mol–1 for the formation of [Co(NH3)4(NO2)(H2O)]2+, and H 2 = (97±1) kJ mol–1 and S 2 = (6±2) JK–1 mol–1 for the formation of [Co(NH3)4(NO2)2]+. Anation of the title complex by N 3 at pH 4.1 also occurs in a stepwise manner ultimately producing thecis-diazido species. At a fixed pH the reaction shows a first-order dependence on [N 3 ] for each step. pH-variation studies at a fixed [N 3 ] show that the hydroxoaqua-form of the complex reactsca. 16 times faster than the diaqua form. Evidence is presented for an ion-pair preequilibrium at high ionic strength (I = 2.0 mol dm–3). Activation parameters obtained from temperature variation studies are: H 1 = (121±1) kJ mol–1 and S 1 = (104±3) JK–1 mol–1 (for the first step anation), and H 2 = (111±2) kJ mol–1 and S 2 = (74±9) JK–1 mol–1 (for the second step anation). The reaction ofcis-tetraaminediaquacobalt(III) ion with salicylate (HSal) has been studied in aqueous acidic medium in the temperature range 39.8–58.2°C. The reaction is biphasic corresponding to the anation of two salicylate ions. The kinetic results for the first phase reaction are compatible with the equation: kobs = kIPQ[HSal]/(1 + Q[HSal]) where Q denotes ion-pair formation constant and kIP is the first-order rate constant for the interchange reaction. The activation parameters obtained from the temperature dependence of rate are: H = (138±3) kJ mol–1 and S = (135±4) JK–1 mol–1. The reaction seems to take place by a dissociative interchange mechanism.  相似文献   

12.
Summary The kinetics of CoIII oxidation of SeIV have been studied in aqueous HClO4. The order with respect to Com is two the order with respect to SeIV is one at low concentrations; two at high concentrations. The latter variation is attributed to the greater reactivity of the SeIV dimier A mechanism involving complexation between oxidant and substrate is proposed. [CoOH]2+ is presumed to be the reactive CoIII species and H2SeO3 and HSeO 3 to be those of SeIV. At 25° C, Ea, H and S for the monomeric path are 125.6±4.0 kJ mol–1, 122.1±3.8 kJ mol–1 and 206±12 JK–1 mol–1 respectively and those for the dimeric path are 88.6±3.6 kJ mol–1, 85.9±3.4 kJ mol–1 and 62.6±11.3 JK–1 mol–1 respectively.  相似文献   

13.
Thermogravimetric (t.g.) and differential scanning calorimetric (d.s.c.) data have been used to study metal–amino acid interactions in adducts of general formula MnCl2 · ngly (gly = glycine, n = 0.7, 2.0, 4.0 and 5.0). All the prepared adducts exhibit only a one step mass loss associated with the release of glycine molecules, except for the 0.7gly adduct, which exhibits two glycine mass loss steps. From d.s.c. data, the enthalpy values associated with the glycine mass loss can be calculated: MnCl2 · 0.7gly = 409 and 399 kJ mol–1, MnCl2 · 2.0gly = 216 kJ mol–1, MnCl2 · 4.0gly = 326 kJ mol–1 and MnCl2 · 5.0gly = 423 kJ mol–1, respectively. The enthalpy associated with the ligand loss, plotted as function of the number of ligands for the n = 2.0, 4.0 and 5.0 adducts, gave a linear correlation, fitting the equation: H (ligand loss)/kJ mol–1 = 67 × (number of ligands, n) + 76. A similar result was achieved when the enthalpy associated with the ligand loss was plotted as a function of the a(COO) bands associated with the coordination through the carboxylate group, 1571, 1575 and 1577 cm–1, respectively, for the n = 2.0, 4.0 and 5.0 adducts, giving the equation H (ligand loss) /kJ mol–1 = 33.5 × a(COO) /cm–1 – 52418.5. This simple equation provides evidence for the enthalpy associated with the ligand loss being very closely related to the electronic density associated with the metal–amino acid bonds.  相似文献   

14.
The kinetics of the interaction of thiourea with [Pt(en)(H2O)2]2+ have been studied spectrophotometrically as a function of [Pt(en)(H2O)2]2+, [thiourea] and temperature at a particular pH(4.0), where the substrate complex exists predominantly as the diaqua species and the thiourea ligand as a neutral molecule. The reaction proceeds via a rapid outer sphere association followed by two slow consecutive steps, the second step exhibiting first order dependence on the aqua ion and thiourea concentrations. The activation parameters for both the steps have been evaluated: (H 1 = 54.8 ± 1.2 kJ mol–1, S 1 = –96 ± 4 J K–1 mol–1, H 2 = 27.9 ± 0.8 kJ mol–1 and, S 2 = –183 ± 2.6 J K–1 mol–1). The low enthalpy of activation and large negative values of entropy of activation indicate an associative mode of activation for both consecutive steps.  相似文献   

15.
Summary The equilibrium vapour pressure of the solid complex, CuCl·MeCN, was measured at several temperatures. The enthalpy and entropy changes according to the following complex formation were determined: CuCl(s)+ MeCN(g)CuCl·MeCN(s); H=–56.9 kJ mol–1 and S=–150 J mol–1 K –1, respectively. Vapour pressure osmometry shows that a monomer-dimer equilibrium of CuCl exists in the acetonitrile solution. The equilibrium constant was found to be 0.93±0.08. A gas chromatographic technique was employed to determine the monomeric species as [Cu(MeCN)4]Cl.  相似文献   

16.
The standard molar enthalpies of formation H f 00B0; (liq) at the temperature t = 298.15 K were determined using combustion calorimetry for N-methyl-3-methyl-3-phenyl-2-butaneamine 1a, N,N-dimethyl-3-methyl-3-phenyl-2-butaneamine 1b N-methyl-2,3-dimethyl-3-phenyl-2-butaneamine 2a, and N,N-dimethyl-2,3-dimethyl-3-phenyl-2-butaneamine 2b. The standard molar enthalpies of vaporization H vap 00B0; of these compounds were obtained from the temperature variation of the vapor pressure measured in a flow system. The following standard molar enthalpies of formation in gaseous phase H f 00B0; (g) are obtained from these data: for 1a – 10.9 ± 1.9; 1b – 3.6 ± 1.8; 1c – 26.6 ± 1.4, and 1d – 23.0 ± 1.8 kJ mol–1. From the standard molar enthalpies of formation for gaseous compounds which are available in the literature, improved values for the increments of the Benson group addivitiy scheme of amines were calculated. They are used to determine the strain enthalpies of the amines 1 and 2 from this investigation.  相似文献   

17.
The kinetics of interaction between DL-Penicillamine and [Rh(H2O)5OH]2+ have been studied spectrophotometrically as a function of [Rh(H2O)5OH2+], [DL-Pen], pH and temperature. The reaction has been monitored at 242 nm, the max of the substituted complex and where the spectral difference between the reactant and product is a maximum. The reaction rate increases with [DL-Pen] and reaches a limiting value at a higher ligand concentration. From the experimental findings an associative interchange mechanism for the substitution process is suggested. The activation parameters (H}=35.8 ± 1.6 kJ mol–1, S=–209 ± 5 J K–1 mol–1) support the proposition. The negative G 0 (–13.6 kJ mol–1) for the first equilibrium step also supports the spontaneous formation of an outersphere association complex.  相似文献   

18.
Summary The oxidation of MeCHO by chromium(VI) has been studied in HClO4 medium over a wide range of experimental conditions and has been found to obey the rate law;v=k[MeCHO][HCrO 4 ][H+]. The calculated H and-S values for the reaction are 30±2kJ mol–1 and 171±7J mol–1deg–1, respectively. The mechanism is discussed in terms of carbon-hydrogen bond cleavage.  相似文献   

19.
Stereochemical nonrigidity of the hexacoordinated (O—Ge)-chelate bis(2-oxo-1-hexahydroazepinylmethyl)dichlorogermane in CDCl3 was studied by dynamic NMR. The activation parameters of the intramolecular rearrangement at the coordination center are G # 298 = 12.3±0.2 kcal mol–1, H # = 16.9±0.2 kcal mol–1, and S # = 15.3±0.7 cal mol–1 K–1. The dissociative mechanism of ligand exchange involving the cleavage of the OGe coordination bond is discussed based on the positive entropy of activation.  相似文献   

20.
Summary Ground state structures and conformational interconversion mechanisms of 25 diaryl compoundsAr 2 Z (Z=CH2, CHR, CH(OH), P-CH3) were analyzed. For tetra(ortho-alkyl)substituted diaryls the cogwheeling mechanism was found as the threshold mechanism. A shift from the cogwheeling mechanism to interconversions via 2-ring flips is found in di(ortho-alkyl)substituted compounds. The ground state structures and interconversion mechanisms of diarylmethylphosphines are very similar to those of the related 1,1-diarylethanes. The interconversion barrier for correlated conrotation of the aryl rings in di(tert-butylphenyl)methanol (20) was measured by low temperature NMR and is in excellent agreement with the calculated value for the 2-ring flipT2 (G (exp.)=48 kJ mol–1; G (calc.)=54 kJ mol–1).
Korrelierte Rotation von Arylringen in Diarylmethyl-, Diarylphosphin- und verwandten Fragmenten. Eine Untersuchung mit Hilfe der empirischen Kraftfeldmethode
Zusammenfassung Die Grundzustandskonformationen und die konformativen Interkonversions-mechanismen von 25 DiarylverbindungenAr 2 Z (Z=CH2, CHR, CH(OH), P-CH3) wurden analysiert. Für tetra(ortho-alkyl)substituierte Diaryle wurde der cogwheeling-Mechanismus als der Interkonversionsmechanismus niedrigster Energie ermittelt. In di(ortho-alkyl)substituierten Verbindungen werden nicht der cogwheeling-Mechanismus sondern 2-ring flips als Interkonversionsmechanismen gefunden. Die Grundzustände und Interkonversionsmechanismen für Diarylmethylphosphine sind sehr ähnlich jenen der verwandten 1,1-Diarylethane. Die Interkonversionsbarriere für die korrelierte Bewegung der Arylringe von Di(tert-butylphenyl)methanol (20) wurde mittels Tieftemperatur-NMR-Spektroskopie ermittelt und ist in sehr guter Übereinstimmung mit dem berechneten Wert für den 2-Ring flipT2 (G (exp.)=48 kJ mol–1; G (calc.)=54 kJ mol–1).
  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