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 734 毫秒
1.
The kinetics of sublimation of bis(2,2,6,6-tetramethyl-3,5-heptanedionato)copper(II), [Cu(tmhd)2] was studied by non-isothermal and isothermal thermogravimetric (TG) methods. The non-isothermal sublimation activation energy values determined following the procedures of Friedman, Kissinger, and Flynn–Wall methods yielded 93 ± 5, 67 ± 2, and 73 ± 4 kJ mol−1, respectively and the isothermal sublimation activation energy was found to be 97 ± 3 kJ mol−1 over the temperature range of 375–435 K. The dynamic TG run proved the complex to be completely volatile and the equilibrium vapor pressure (pe)T of the complex over the temperature range of 375–435 K determined by a TG-based transpiration technique, yielded a value of 96 ± 2 kJ mol−1 for its standard enthalpy of sublimation (ΔsubH°).  相似文献   

2.
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   

3.
Solid complexes of lighter lanthanide nitrates with N,N′-dinaphthyl-N,N′-diphenyl-3,6-dioxaoctanediamide (DDD), Ln(NO3)3(DDD) (Ln = La---Nd, Sm) have been prepared in non-aqueous media. These complexes have been characterized by elemental analysis, conductivity measurements, IR spectra, electronic spectra and TG-DTA techniques. In all the complexes, DDD and NO3 are coordinated to the lanthanide ions as tetradentate and bidentate ligands, respectively. The differences in the IR and electronic spectra between these complexes and lanthanide nitrate complexes with N,N,N′,N′-tetraphenyl-3,6-dioxaoctanediamide (TDD) are discussed.  相似文献   

4.
A practical procedure suitable for large scale lithiation of N-tosylindoles and subsequent addition to ketones is described. Bis(N,N-dimethylaminoethyl) ether was found to stabilize 2-lithio-N-tosylindole 1A at −25 °C [The temperatures cited are internal temperatures unless otherwise stated]. Addition of this reagent allows the lithiation of N-tosyl indoles and subsequent addition to ketones to operate at −25 °C, a temperature suitable for large scale reactions.  相似文献   

5.
Differential scanning calorimetry and high temperature oxide melt solution calorimetry are used to study enthalpy of phase transition and enthalpies of formation of Cu2P2O7 and Cu3(P2O6OH)2. α-Cu2P2O7 is reversibly transformed to β-Cu2P2O7 at 338–363 K with an enthalpy of phase transition of 0.15 ± 0.03 kJ mol−1. Enthalpies of formation from oxides of α-Cu2P2O7 and Cu3(P2O6OH)2 are −279.0 ± 1.4 kJ mol−1 and −538.8 ± 2.7 kJ mol−1, and their standard enthalpies of formation (enthalpy of formation from elements) are −2096.1 ± 4.3 kJ mol−1 and −4302.7 ± 6.7 kJ mol−1, respectively. The presence of hydrogen in diphosphate groups changes the geometry of Cu(II) and decreases acid–base interaction between oxide components in Cu3(P2O6OH)2, thus decreasing its thermodynamic stability.  相似文献   

6.
Adsorption (at a low temperature) of nitrogen on the protonic zeolite H-Y results in hydrogen bonding of the adsorbed N2 molecules with the zeolite Si(OH)Al Brønsted-acid groups. This hydrogen-bonding interaction leads to activation, in the infrared, of the fundamental N–N stretching mode, which appears at 2334 cm−1. From infrared spectra taken over a temperature range, the standard enthalpy of formation of the OH···N2 complex was found to be ΔH0 = −15.7(±1) kJ mol−1. Similarly, variable-temperature infrared spectroscopy was used to determine the standard enthalpy change involved in formation of H-bonded CO complexes for CO adsorbed on the zeolites H-ZSM-5 and H-FER; the corresponding values of ΔH0 were found to be −29.4(±1) and −28.4(±1) kJ mol−1, respectively. The whole set of results was analysed in the context of other relevant data available in the literature.  相似文献   

7.
The reaction of Zn(acac)2 with btaH (1,2,3-benzotriazole) in dmf yielded the pentanuclear complex [Zn5(bta)6(acac)4(dmf)]·dmf (1·dmf). In the presence of pyrazine, the pentanuclear [Zn5(bta)6(acac)4(dmf)]·3.7dmf (2·3.7dmf) and enneanuclear [Zn9(bta)12(acac)6]·6dmf (3·6dmf) complexes were formed, whereas in the presence of 4,4′-bpy the 1D coordination polymer [Zn(acac)2(4,4′-bpy)]n (4) was isolated. The molecular structures of 1·dmf and 2·3.7dmf reveal that the [Zn5] clusters consist of four ZnII ions which span the corners of a tetrahedron and the fifth resides at its centre. The molecular structure of 3·6dmf reveals that the [Zn9] clusters consist of two corner sharing tetrahedra and the structure can be described as the addition of two [Zn5] clusters of 1·dmf and/or 2·3.7dmf followed by the simultaneous abstraction of [Zn(acac)2] and dmf molecules; this alternative was accomplished by recrystallization of 1·dmf from dmf which yielded 3·6dmf. Each of the μ3N:κN′:κN′′ benzotriazolate ligands in 1·dmf, 2·3.7dmf and 3·6dmf spans an edge of the tetrahedron. The molecular structure of 4 reveals mononuclear [Zn(acac)2] units bridged via 4,4′-bpy molecules to 1D coordination polymer. Characteristic IR bands of the four complexes are discussed in terms of the coordination modes of the ligands and known structures.  相似文献   

8.
Mixed-chelate complexes of ruthenium have been synthesized using tridentate Schiff-base ligands (TDLs) derived from condensation of 2-aminophenol or 2-aminobenzoic acid with aldehydes (salicyldehyde, 2-pyridinecarboxaldehyde), and tmeda (tetramethylethylenediamine). [RuIII(hpsd)(tmeda)(H2O)]+ (1), [RuIII(hppc)(tmeda)(H2O)]2+ (2), [RuIII(cpsd)(tmeda)(H2O)]+ (3) and [RuIII(cppc)(tmeda)(H2O)]2+ (4) complexes (where hpsd2− = N-(hydroxyphenyl)salicylaldiminato); hppc = N-(2-hydroxyphenylpyridine-2-carboxaldiminato); cpsd2− = (N-(2-carboxyphenyl)salicylaldiminato); cppc = N-2-carboxyphenylpyridine-2-carboxaldiminato) were characterized by microanalysis, spectral (IR and UV–vis), conductance, magnetic moment and electrochemical studies. Complexes 14 catalyzed the epoxidation of cyclohexene, styrene, 4-chlorostyrene, 4-methylstyrene, 4-methoxystyrene, 4-nitrostyrene, cis- and trans-stilbenes effectively at ambient temperature using tert-butylhydroperoxide (t-BuOOH) as terminal oxidant. On the basis of Hammett correlation (log krel vs. σ+) and product analysis, a mechanism involving intermediacy of a [Ru–O–OBut] radicaloid species is proposed for the catalytic epoxidation process.  相似文献   

9.
The synthesis of a new series of six-membered N,N′-diarylsubstituted methylene-bis-dihydro-2H-1,3-benzoxazines (5a-e) was achieved in excellent yields by Mannich-type condensation of N,N′-diarylsubstituted methylene-bis-o-hydroxybenzyl amines (4a- e) with formaldehyde in chloroform at reflux. These amines (4a-e) were obtained by the reduction of N, Nr-diarylsubstituted methylene-bis-o-hydroxybenzyl imines (3a-e) with NaBH4, which inturn were obtained by the condensation of methylene-bissalicylaldehyde (2) with various substituted arylamines.  相似文献   

10.
The novel triethylantimony(v) o-amidophenolato (AP-R)SbEt3 (R = i-Pr, 1; R = Me, 2) and catecholato (3,6-DBCat)SbEt3 (3) complexes have been synthesized and characterized by IR, NMR spectroscopy (AP-R is 4,6-di-tert-butyl-N-(2,6-dialkylphenyl)-o-amidophenolate, alkyl = isopropyl (1) or methyl (2); 3,6-DBCat is 3,6-di-tert-butyl-catecholate). Complexes 13 have been obtained by the oxidative addition of corresponding o-iminobenzoquinones or o-benzoquinones to Et3Sb. The addition of 4,6-di-tert-butyl-N-(3,5-di-tert-butyl-2-hydroxyphenyl)-o-iminobenzoquinone to Et3Sb at low temperature gives hexacoordinate [(AP-AP)H]SbEt3 (4) which decomposes slowly in vacuum with the liberation of ethane yielding pentacoordinate complex [(AP-AP)]SbEt2 (5). [(AP-AP)H]2− is O,N,O′-tridentate amino-bis-(3,5-di-tert-butyl-phenolate-2-yl) dianion and [(AP-AP)]3− is amido-bis-(3,5-di-tert-butyl-phenolate-2-yl) trianion. The decomposition of 45 accelerates in the presence of air. o-Amidophenolates 1 and 2 bind molecular oxygen to give spiroendoperoxides Et3Sb[L-iPr]O2 (6) or Et3Sb[L-Me]O2 (7) containing trioxastibolane rings. This reaction proceeds slowly and reaches the equilibrium at 15–20% conversion five times more than for (AP-R)SbPh3 analogues. Molecular structures of 1 and 5 were determined by X-ray analysis.  相似文献   

11.
A DFT computational study is performed on different Cp2TiIV(L,L′-BID) complexes with L,L′-BID = dioxolene, dithiolene or diselenolene. A fragment analysis of the titanocene-ligand bonding in the DFT optimized geometries showed that out of plane folding for maximum Ti ← L π donation increases Cp2TiIV(O,O′-BID) (35°) < Cp2TiIV(S,S′-BID) (43–49°) < Cp2TiIV(Se,Se′-BID) (48–53°).  相似文献   

12.
The electrochemical reduction of the black dye photosensitizer [(H3-tctpy)RuII(NCS)3] (H3-tctpy=2,2′:6′,2′′-terpyridine-4,4′,4′′-tricarboxylic acid) used in photovoltaic cells has been found to be a complex process when studied in dimethylformamide. At low temperatures, fast scan rates and at a glassy carbon electrode, the chemically reversible ligand based one-electron reduction process [(H3-tctpy)Ru(NCS)3]+e[(H3-tctpy√)Ru(NCS)3]2− is detected. This process has a reversible half-wave potential (Er1/2) of −1585±20 mV versus Fc/Fc+ at 25°C. Under other conditions, a deprotonation reaction occurs upon reduction, which produces [(H3−x-tctpyx)Ru(NCS)3](1+x)− and hydrogen gas. Mechanistic pathways giving rise to the final products are discussed. The Er1/2-value for the ligand based reductions of the deprotonated complex is 0.70 V more negative than for [(H3-tctpy)Ru(NCS)3]. Consequently, data obtained from molecular orbital calculations are consistent with the reaction [(H3-tctpy)Ru(NCS)3]+e→[(H2-tctpy)Ru(NCS)3]2−+1/2H2 yielding the monodeprotonated complex as the major product obtained after electrochemical reduction of [(H3-tctpy)Ru(NCS)3]. The Er1/2-values for the metal based RuII/III process differ by 0.30 V when data obtained for the protonated and deprotonated forms of the black dye are compared. Electronic spectra obtained during the course of experiments in an optically transparent thin layer electrolysis configuration are consistent with the overall reaction scheme proposed on the basis of voltammetric measurements and molecular orbital calculations. Reduction studies on the free ligand, H3-tcpy, are consistent with results obtained with [(H3-tctpy)Ru(NCS)3].  相似文献   

13.
The standard (p = 0.1 MPa) molar enthalpies of formation of 2-, 3- and 4-cyanobenzoic acids were derived from their standard molar energies of combustion, in oxygen, at T = 298.15 K, measured by static bomb combustion calorimetry. The Calvet high temperature vacuum sublimation technique was used to measure the enthalpies of sublimation of 2- and 3-cyanobenzoic acids. The standard molar enthalpies of formation of the three compounds, in the gaseous phase, at T = 298.15 K, have been derived from the corresponding standard molar enthalpies of formation in the condensed phase and standard molar enthalpies for phase transition. The results obtained are −(150.7 ± 2.0) kJ · mol−1, −(153.6 ± 1.7) kJ · mol−1 and −(157.1 ± 1.4) kJ · mol−1 for 2-cyano, 3-cyano and 4-cyanobenzoic acids, respectively. Standard molar enthalpies of formation were also estimated by employing two different methodologies: one based on the Cox scheme and the other one based on several different computational approaches. The calculated values show a good agreement with the experimental values obtained in this work.  相似文献   

14.
Thermophysical and thermochemical studies have been carried out for crystalline parabanic acid. The thermophysical study was made by differential scanning calorimetry, DSC, over the temperature interval between T = (263 and 473) K. Two phase transitions were found: at T = (392.3 ± 1.6) K with the enthalpy of transition of (2.1 ± 0.4) kJ · mol−1 and at T = (509.8 ± 1.5) K, when the compound was scanned to its fusion temperature. The standard (p = 0.1 MPa) molar enthalpy of formation, at T = 298.15 K, for crystalline parabanic acid was determined using static-bomb combustion calorimetry as −(590.2 ± 1.0) kJ · mol−1. The standard molar enthalpy of sublimation, at T = 298.15 K, was derived from the variation of their vapour pressures, measured by the Knudsen-effusion method, with the temperature. These two thermochemical parameters yielded the standard molar enthalpy of formation in the gaseous phase, at T = 298.15 K, as −(470.8 ± 1.2) kJ · mol−1.  相似文献   

15.
Solvothermal reaction assisted with microwave leads to the formation of two unique heterometallic cubic clusters [Ni3M′(L)3(OH)(CH3CN)3]2·CH3CN (M′=K for 1 and M′=Na for 2, where L is an anion of 2-[(2-hydroxy-3-methoxy-benzylidene)-amino]-ethanesulfonate) with higher efficiency, yields and purity than those without it. The 6-metallacrown-3 [Ni3(OH)(L)3] groups exhibit interesting ion trapping and self-assembly of size-different Na+ and K+ through form recognition and coordination activity in 1 and 2. The magnetic studies for 1 and 2 suggest that the {Ni3M′O4} (M′=K and Na) cores both display dominant ferromagnetic interactions from the nature of the binding modes of μ3-O (oxidophenyl) and μ3-OH.  相似文献   

16.
A series of complexes of the type M(Phca2en)X2, where Phca2en=N,N′-bis(β-phenyl-cinnamaldehyde)-1,2-diiminoethane, M(II)=Co, Ni or Zn and X=Cl, Br, I or NCS have been synthesized and characterized. The crystal and molecular structures of Co(Phca2en)Cl2 (2), Ni(Phca2en)Br2 (5) and Zn(Phca2en)Cl2 (6) were determined by X-ray crystallography from single-crystal data. Complexes 2 and 5 are isomorph and isostructure, in which the coordination polyhedron about the central metal ion is distorted tetrahedron with Cl---Co---Cl, 110.17(6)°; N---Co---N, 84.16(13)° and Cl---Zn---Cl, 112.02(6)°; N---Zn---N, 83.45(16)°. The complex 5 crystallizes in triclinic system with two molecules per asymmetric unit, both having nickel ion in distorted tetrahedral geometry, Br---Ni---Br, 122.645(18)° and 125.729(18)°; N---Ni---N, 84.63(9)° and 85.08(9)°. These structures consist of intermolecular hydrogen bonds of the type C---HX. The formation of the C---HM weak intramolecular hydrogen bonds due to the trapping of C---H bonds in the vicinity of the metal atoms are reported for 2, 5 and 6. A 1H NMR study of Zn complexes gives further evidence for the presence of such interactions and their significance. The spectral properties of the above complexes are also discussed.  相似文献   

17.
Electrochemical and spectroscopic (EPR, UV–Vis, IR) studies of the aromatic secondary amines N,N′-diphenyl-1,4-phenylenediamine (DPPD), N-phenyl-N′-isopropyl-p-phenylene diamine (IPPD), N-phenyl-N′-(α-methylbenzyl)-p-phenylenediamine (SPPD) and N-phenyl-N′-(1,3-dimethyl-butyl)-p-phenylenediamine (6PPD), which represent the most important group of antioxidants used in the rubber industry, are presented. During oxidation, all the compounds show reversible redox couples in acetonitrile/0.1 M TBABF4. The first oxidation potential depends substantially on the R substituent at the –N′H– moiety. Very similar UV–VIS spectra of monocation radicals and dications for all the compounds were observed by applying anodic oxidation as well as oxidation by tert-butyl hydroperoxide both in air and in inert atmosphere. The samples with N′-bonded aliphatic carbon in the molecule (e.g. IPPD) heated in air undergo consecutive chemical reactions leading to the formation of –N′C– group. By the use of RO2 radicals only very low concentration of nitroxide radicals was obtained. Very high concentration of nitroxide radicals was achieved using 3-chloroperbenzoic acid. In the oxidation of investigated aromatic secondary amines with powder PbO2 no EPR spectra were observed and UV–Vis and IR studies indicate the rapid formation of the final dehydrogenated oxidation product.  相似文献   

18.
Based on the ligand dppz (dppz = dipyrido-[3,2-a:2′,3′-c]phenazine), a new ligand pbtp (pbtp = 4,5,9,11,14-pentaaza-benzo[b]triphenylene) and its polypyridyl ruthenium(II) complexes [Ru(phen)2(pbtp)]2+ (1) (phen = 1,10-phenanthroline and [Ru(bpy)2(pbtp)]2+ (2) (bpy = 2,2′-bipyridine) have been synthesized and characterized by elemental analysis, ES-MS and 1H NMR spectroscopy. The DNA-binding of these complexes were investigated by spectroscopic methods and viscosity measurements. The experimental results indicate that both complexes 1 and 2 bind to CT-DNA in classical intercalation mode, and can enantioselectively interact with CT-DNA. It is interesting to note that the pbtp ruthenium(II) complexes, in contrast to the analogous dppz complexes, do not show fluorescent behavior when intercalated into DNA. When irradiated at 365 nm, both complexes promote the photocleavage of pBR 322 DNA.  相似文献   

19.
The combustion energies for 2-acetylpyrrole (cr) and 2-acetylfuran (cr) were determined using a static bomb calorimeter, whereas the combustion energy of 2-acetylthiophene (l) was determined with a rotating bomb calorimeter; both calorimeters have been recently described. The molar combustion energies obtained were: −(3196.1 ± 0.6) kJ mol−1 for 2-acetylpyrrole, −(2933.8 ± 0.7) kJ mol−1 for 2-acetylfuran, and −(3690.4 ± 0.8) kJ mol−1 for 2-acetylthiophene. From these combustion energy values, the standard molar enthalpies of formation in the condensate phase were obtained as: −(163.51 ± 0.97) kJ mol−1, −(283.50 ± 1.06) kJ mol−1 and −(123.93 ± 1.15) kJ mol−1, respectively. The obtained values of combustion and formation enthalpies of 2-acetylthiophene are in concordance with the reported previously. For the two last compounds, polyethene bags were used as an auxiliary material in the combustion experiments. The heat capacities and purities of the compounds were determined using a differential scanning calorimeter.  相似文献   

20.
The new salt, tetra-n-butylammonium bis(benzene-1,2-dithiolato(2−)-κ2S,S′)platinate(III), [NBu4][Pt(C6H4S2)2] (1), has been synthesized in ethanol/water, and fully characterized by single crystal X-ray structure determination. The central platinum in the complex ion [Pt(bdt)2] is tetracoordinated by the S atoms of the bdt2− ligands (bdt2− is benzene-1,2-dithiolate) in a square-planar geometry. The well-resolved frozen solution EPR spectrum exhibits rhombic symmetry. The room temperature effective magnetic moment (μeff = 1.80 Bohr magneton) is in line with this spectrum and strongly supports the Pt(III) oxidation state in 1. This observation is in excellent agreement with previous results reported on closely related Ni(III), Pd(III) and Pt(III) species.  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