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 609 毫秒
1.
The possibility of initiating detonation in the supersonic flow of the H2 + O2 (air) mixture behind the front of the inclined shock wave by O2 molecule excitation to the O2(a 1g) and O2(b 1·+ g) states by laser radiation with a wavelength I = 1.268 m and 762 nm is considered. Resonance laser radiation intensifies chain combustion due to the formation of new pathways for generating active atoms O· and H· and radicals OH and has a substantially nonthermal character. Even at low (3 kJ/cm2) energies of radiation with I = 762 nm applied to the gas, detonation combustion can occur even at a distance of 1 m from the front at the gas temperature as high as 600 K.  相似文献   

2.
The solubility, solubility product and the thermodynamic functions for the CeF3–H2O system have been measured using the radiometric, conductometric and potentiometric techniques. The radiometric values for the solubility and solubility product, the lowest and more acceptable for reasons cited in previous papers, are 3.14·10–5 M and 2.17·10–17 respectively. The enthalpy change measured by the conductometric method is almost twice as that obtained by potentiometric method due to abnormal conductances registered at higher temperatures. The average values for Ho and Go and So at 298 K are 53.0±17.4, 91.7±4.0 and –129.7±58.2 KJ·mol–1 respectively. The positive values for Ho and Go and the negative value for So are indicative of the low solubility of this salt in water. The stability constants for the mono- and difluoride complexes of Ce(III) have been determined potentiometrically using unsaturated solution mixtures of Ce(III) and F. These values for CeF+ and CeF 2 + are 997±98 and (1.03±0.44)·105, respectively. Studies on pH dependence of the solubility shows that the solubility reaches a minimum value at a pH of about 3.2.  相似文献   

3.
Numerous experimental data for the cyclization of free radicals C·H2(CH2)nCH=CH2 cyclo-[(CH2)n+1CH(C·H2)], and C·H2(CH2)nCH=CHR cyclo-[(CH2)n+1C·HCHR] were analyzed in the framework of the parabolic model. The activation energy of thermoneutral (H e = 0) cyclization E e0 decreases linearly with an increase in the energy of cycle strain E rsc: E e0(n) (kJ mol–1) = 85.5 – 0.44E rsc(n) (n is the number of atoms in the cycle). The activation entropy of cyclization S # also depends on the cycle size: the larger the cycle, the lower S #. A linear dependence of S # on the difference between the entropies of formation S° of cyclic hydrocarbon and the corresponding paraffin was found: S # = 1.00[S°(cycle) – S°(CnH2n+2)]. The E e0 values coincide for cyclization reactions with the formation of the six-membered cycle and the bimolecular addition of alkyl radicals to olefins.  相似文献   

4.
The pressure dependence of excited-state proton transfer equilibria has been examined for aqueous solutions of several substituted napthalene dyes, in particular 1-dimethylaminonaphthalene-5-sulfonic acid (DANS). The pressure-induced shift in equilibrium is characterized by volume changes spanning the range V *=–18 cm3 mole–1 to V *=+4 cm3-mole–1. A deuterium oxide solvent isotope effect is evident in the pressure response of DANS, leading to a 35% smaller V* in D2O relative to H2O.  相似文献   

5.
4-Arylurazoles are selectively oxidized with Fe2(NO3)3 · 9H2O to the corresponding 1-1,2,4-triazoline-3,5-diones.  相似文献   

6.
The complex of [Tb2(p-MOBA)6(PHEN)2] (p-MOBA=C8H7O3,p-methoxybenzoate; PHEN=C12H8N2, 1,10-phenanthroline) was prepared and characterized by elemental analysis and IR spectroscopy. The thermal behavior of Tb2(p-MOBA)6(PHEN)2 in a static air atmosphere was investigated by TG-DTG, DTA, SEM and IR techniques. By the kinetic method of processing thermal analysis data put forward by Malek et al., it is defined that the kinetic model for the first-step thermal decomposition is SB(m,n). The activation energy E for this step reaction is 140.92 kJ mol-1, the enthalpy of activation H is 136.06 kJ mol-1, the Gibbs free energy of activation G is 145.16 kJ mol-1, the entropy of activation S is -15.53 J mol-1, and the pre-exponential factor lnA is 29.26. The lifetime equation at mass loss of 10% was deduced as ln =-28.72+1.943·104/T by isothermal thermogravimetric analysis.  相似文献   

7.
The melting process of NC is studied by using modulated differential scanning calorimetry (MDSC) technique, the microscope carrier method for measuring the melting point and the simultaneous device of the solid reaction cell in situ/RSFT-IR. The results show that the endothermic process in the MDSC curve is reversible. It is caused by the phase change from solid to liquid of the mixture of initial NC, decomposition partly into condensed phase products. The values of the melting point, melting enthalpy (Hm), melting entropy (Sm), the enthalpy of decomposition (Hdec) and the heat-temperature quotient (Sdec) obtained by the MDSC curve of NC at a heating rate of 10 K min–1 are 476.84 K, 205.6 J g–1, 0.4312 J g–1 K–1, –2475.0 J g–1 and –5.242 Jg–1K–1, respectively. The MDSC results of NC with different nitrogen contents show that with increasing the nitrogen content in NC, the absolute values of Hm, Sm, Hdec and Sdec increase.This revised version was published online in November 2005 with corrections to the Cover Date.  相似文献   

8.
The enthalpies of formation (H f°) of 16 alkyl-substituted phenyl and benzyl radicals (R·) were determined for the first time by the published values of energies of R—X bond dissociation. For the initial molecules of RX, alkyl-substituted benzenes, the additive-group procedure was developed for the calculation of H f°. In the framework of the additive-group model for considered R·, we studied the structure-property interrelation, analyzed the obtained H f°(R·) values, and confirmed their reliability. The influence of nonvalent interactions on H f°(R·) was systematized and detailed. The parameters, from which it is possible to calculate H f° of the 51 radicals, were proposed.  相似文献   

9.
Summary The experimental data available on the thermodynamic functions 0 forMnHF hydrofluorides [M=Li, Na, K, Rb, Cs, NH4, Ag(I) and Tl(I);n=1–3] have been evaluated additively. The unknown values of 0 forn=0÷7 are predicted.
Additive Voraussagen der thermodynamischen Funktionen von Hydrogenfluoriden (Kurze Mitt.)
Zusammenfassung Die vorhandenen experimentellen Daten über die thermodynamischen Funktionen 0 von HydrogenfluoridenMnHF [M=Li, Na, K, Rb, Cs, NH4, Ag(I) und Tl(I);n=1–3] werden linear ausgeglichen und die fehlendenden Werte für 0 mitn=0÷7 vorausgesagt.
  相似文献   

10.
The thermal dissociation processes of clathrates [CuPy4(NO3)2]·2G (G=tetrahydrofuran, chloroform) were studied. Thermodynamic parameters (H AV 0,S 298 0 and G 298 0) of the processes, producing solid host complex [CuPy4(NO3)2] and gaseous guest (G), were determined from the strain measurements. These data are compared with previous data for clathrates of the host complex with benzene and pyridine. Quasi-equilibrium thermogravimetry was used to investigate the step-by-step character of the dissociation processes. Kinetic studies were carried out for clathrate [CuPy4(NO3)2]·2THF.This revised version was published online in November 2005 with corrections to the Cover Date.  相似文献   

11.
Crystallization of 2RbBr · MnBr2 · 2H2O, the only double salt obtained under standard conditions from saturated aqueous rubidium–manganese bromide solutions, was theoretically predicted using the hard and soft Lewis acids and bases concept and Pauling's rules. The RbBr—MnBr2—H2O system was thermodynamically simulated by the Pitzer model assuming a solubility diagram of three branches only: RbBr, 2RbBr · MnBr2 · 2H2O and MnBr2 · 4H2O. The theoretical result was experimentally proved at 25°C by the physicochemical analysis method and formation of the new double salt 2RbBr · MnBr2 · 2H2O was established. It was found to crystallize in a triclinic crystal system, space group –P1, a = 5.890(1) Å, b = 6.885(1) Å, c = 7.367(2) Å, = 66.01(1)°, = 87.78(2)°, = 84.93(2)°, V = 271.8(1) Å3, Z = 1, D x = 3.552 g-cm–3. The binary and ternary ion interaction parameters were calculated and the solubility isotherm was plotted. The standard molar Gibbs energy of the synthesis reaction, rG m o , of the double salt 2RbBr · MnBr2 · 2H2O from the corresponding simple salts RbBr and MnBr2 · 4H2O, as well as the standard molar Gibbs energy of formation, fG m o , and standard molar enthalpy of formation fH m o of the simple and double salts were calculated.  相似文献   

12.
Thermodynamic parameters (H ex 0 and S ex 0 ) for the overall extractions of monovalent metal (Na, K, Rb, and Tl) picrates with benzo-18-crown-6 (B18C6), and those (H D,L 0 and S D,L 0 ) for the distribution of B18C6 were determined between chloroform and water. All the extracted B18C6 complexes were l:1:1 complexes (B18C6:metal ion: picrate anion). The H ex 0 and S ex 0 values for all the metals are negative. Every extraction of the metal picrate with B18C6 is completely enthalpy driven. The H D,L 0 and S D,L 0 values of B18C6 are both positive, and the partition of B18C6 is entirely entropy driven. Enthalpy (H ex,ip 0 ) and entropy changes (S ex,ip 0 ) for ion-pair extractions of B18C6-metal ion complexes with picrate anions were calculated. All the H ex,ip 0 and S ex,ip 0 values are negative, and the ion-pair extractions are completely enthalpy driven.  相似文献   

13.
Complexes of formula CuCl2 · 2arg and CuCl2 · 4ala (arg = arginine; ala = alanine) were prepared at room temperature by a solid state route. The metal–amino acid solid state interactions were studied by i.r. spectroscopy and solution calorimetry. For both complexes, participation of the carboxylate group as well as nitrogen in coordination are inferred, based on the i.r. data. For the copper–arginine compound, the calculated thermochemical parameters are: rHm = –114.9 ± 1.42 and fHm = –1608.3 ± 11.6 kJ mol–1. For copper–alanine compound, a complete set of thermochemical parameters were calculated: rHm = –18.0 ± 0.9; fHm = –2490.4 ± 4.3; DHm = 597.2 ± 17.7; MHm = 771.9 ± 18.7; gHm = 627.1 ± 22.3 and D (Cu–L) = 156.8 ± 5.7 kJ mol–1. Based on rHm and dissolution enthalpy values, a stronger intermolecular solid state interaction can be inferred for the arginine complex, than for the alanine one complex, probably due to the formation of intermolecular hydrogen bonds in the former.  相似文献   

14.
Zusammenfassung Imidazolin-3-thione-(5) sind zyklische Thioamide, die auch als Thiolimide (5-Mercapto-2H-imidazole) reagieren können. Die Reaktionsmöglichkeiten des exozyklisch gebundenen S-Atoms werden am Beispiel des 2-Methyl1-2,4-diphenyl-imidazolin-3-thion-(5) demonstriert. Vom Thiolimid-Tautomeren leiten sich Cd- und Cu-Salze, S-Alkyl- und S-Acylderivate sowie das Bis-[2H-imidazol-5-yl]-disulfid ab. Aus dem Thioamid entsteht mit H2O2 in Methanol das S-Oxid, das sich durch Acetylierung und nachfolgende Hydrolyse in das Disulfid überführen läßt. In höherer Ausbeute entsteht das Disulfid direkt aus dem Imidazolin-3-thion-(5) durch Behandeln mitAc 2O und H2O2 in Gegenwart von Pyridin. Disulfid oder Imidazolin-3-thion-(5) ergeben mit SO2Cl2 5-Chlor-2H-imidazol, welches mit Aminen zu 5-Amino-2H-imidazolen ungesetzt werden kann. Aus der reduktiven Entschwefelung von Imidazolin-3-thion mit LiAlH4 resultieren unter Ringöffnung Gemische zweier isomerer Diamine.
3-Imidazoline-5-thiones are cyclic thioamides which can react also as thiol imides (5-mercapto-2H-imidazoles). The reactions of the exocyclic S atom are demonstrated with 2-methyl-2.4-diphenyl-3-imidazoline-5-thione.Derivatives of the tautomer thiolimide are cadmium and copper salts, S-alkyl and S-acyl derivatives and the bis-[2H-imidazol-5-yl]-disulphide. Oxidation of the thioamide in methanolic solution with H2O2 gives the corresponding disulphide. Better yields of the disulphide are obtained when 3-imidazoline-5-thione is treated simultaneously withAc 2O, H2O2 and pyridine. Reaction of the disulphide or of the 3-imidazoline-5-thione with SO2Cl2 gives 5-chloro-2H-imidazole, which can be converted to 5-amino-2H-imidazoles with amines. Desulphurization of the 3-imidazoline-5-thione under reducing conditions (LiAlH4) causes ring opening to yield a mixture of two isomeric diamines.


Teil der DissertationA. Wegerhoff, T. H. Aachen (1964).

Teil der DissertationG. Kriebel, T. H. Aachen (1965).  相似文献   

15.
The present work aims chiefly to study the thermal behaviour of complex compounds with general formula: [M(HL)xH2O](A)yH2O (where HL=C13H11N4O2=6-(2-pyridylazo)-3-acetamidophenol (PAAP), M=Cu(II), Zn(II), Cd(II) and Fe(III) x=1, 3; y=2, 5) while A=CH3COO (Ac), Cl2. The second formula is [M(H2L)xH2O]Cl2yH2O, (where H 2 L=C13H12N4O2 (PAAP), M=Ni(II), Co(II) x=3; y=4, 6). The compounds were identified by elemental analysis, FT-IR spectra and TG/DTG,DTA methods. It was found that during the thermal decomposition of complex compounds water molecules of crystallization are released in the first step. In the next step the pyrolysis of organic ligand takes place. Metal oxide remained as a solid product of the thermal decomposition. Mass spectroscopy has been used for the determination of the thermal decomposition on the intermediate products. It was found that the thermal stability of the studied compounds increases as the ionic radii decreases. The activation energy E, the entropy change S *, the enthalpy H * change and Gibbs free energy change G * were calculated from TG curve.This revised version was published online in November 2005 with corrections to the Cover Date.  相似文献   

16.
Zusammenfassung Die Gleichgewichtsselektivitätskoeffizienten wurden als Funktion der Temperatur für den Austausch von n-Alkylsulfaten ROSO3(Na) mit der Chloridform des stark basischen Anionenaustauscherharzes AG-IX8 bestimmt. Selektivitätskoeffizienten und thermodynamische Größen (G, H, S) wurden für die Austausche in den homologen Reihen der Anionen ROSO3 und RSO3 verglichen. Die Austauschentropien nahmen mit zunehmender Kettenlänge für beide Ionenfamilien in ähnlicher Weise zu; das Bestehen eines allgemeinen hydrophoben Selektivitätseffektes — wirksam für alle gelösten Ionen, die gesättigte Kohlenwasserstoffketten enthalten — wurde festgestellt.
Anion exchange selectivity sequences in homologous series of hydrophobic ions
The equilibrium selectivity coefficients as a function of temperature were determined for the exchange of n-alkylsulfates ROSO3(Na) with the chloride form of the strongly basic anion exchange resin AG-IX8. Selectivity coefficients and thermodynamic quantities (G, H, S) were compared for the exchanges involving the homologous series of ROSO3 and RSO3 anions. The entropies of exchange increased with increasing chain length in a similar fashion for both families of ions and the existence of a general hydrophobic selectivity effect active for all ionic solutes containing saturated hydrocarbon chains was established.


Mit 3 Abbildungen

Herrn Professor Dr.H. Nowotny gewidmet.  相似文献   

17.
Complexation of sulfones (S) with the -diketonate Eu(Fod)3 (Fod–heptafluorodimethyloctanedione) in the ground and excited electronic states in benzene solutions was studied. The stability constants and thermodynamic parameters for the formation of complexes Eu(Fod)3 · S in the ground state (K, H 0, S 0) and Eu(Fod)3 * · S in the excited state (K*, H 0 *, S 0 *) were determined. The excitation of ff transitions of Eu(III) was found to enhance the stability of Eu(Fod)3 · S complexes, apparently due to an increase in the acceptor ability of the Eu(III) chelate. This fact confirms the involvement of the 4f orbital in the chemical bond formation. The compensation effect was observed for the thermodynamic parameters: S 0 = (2.9 ± 0.3) × 10–3H 0 + (35.0 ± 4.0) in the ground and S 0 * = (3.3 ± 0.3) × 10–3H 0 * + (49.0 ± 5.0) in the excited states of Eu(Fod)3. It was shown that electronic excitation of the 4f orbital of Eu(Fod)3 influences isotopic effects in complexation with sulfolanes.  相似文献   

18.
Recent determination of the standard enthalpy of formation of the ammonium azide f H O (NH N 3,c) and the assignment of the viscosity B-coefficient for the azide anion, B(N 3 ,aq), in aqueous solution enable us to estimate the standard enthalpy of formation of the gaseous azide anion, f H O (N 3 –,g , — a thermochemical magnitude in some dispute — to be 192 kJ-mol–1.  相似文献   

19.
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   

20.
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   

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