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 781 毫秒
1.
A systematic thermodynamic and kinetic study of the entire SFxCl (x = 0-5) series has been carried out. High-level quantum chemical composite methods have been employed to derive enthalpy of formation values from calculated atomization and isodesmic energies. The resulting values for the SCl, SFCl, SF2Cl(C1), SF3Cl(Cs), SF4Cl(Cs) and SF5Cl molecules are 28.0, −36.0, −64.2, −134.3, −158.2 and −237.1 kcal mol−1. A comparison with previous experimental and theoretical values is presented. Statistical adiabatic channel model/classical trajectory, SACM/CT, calculations of selected complex-forming and recombination reactions of F and Cl atoms with radicals of the series have been performed between 200 and 500 K. The reported rate coefficients span over the normal range of about 6 × 10−12 and 5 × 10−11 cm3 molecule−1 s−1 expected for this type of barrierless reactions.  相似文献   

2.
Methyl(4-methoxyphenyl)(2,2-bipyridine)palladium(II) (1) reacts with bis(4-chlorophenyl) diselenide in dichloromethane to form an equilibrium with the Pd(IV) complex Pd(SeC6H4Cl)2Me(C6H4OMe)(bpy) (2) for which the forward reaction exhibits ΔH=−130±12 kJ mol−1 and ΔS=−472±49 J K−1 mol−1, and with K=754±145 at −25 °C. The Pd(IV) complex is isolable at −40 °C, and when the equilibrium mixture is kept at −25 °C, a temperature at which the Pd(II) complex is stable, selective reductive elimination of Me-SeC6H4Cl occurs very slowly from the Pd(IV) complex to form Pd(SeC6H4Cl)(C6H4OMe)(bpy) (3). In contrast, (ClC6H4Se)2 reacts with PdMe2(dmpe) (4) [dmpe=1,2-bis(dimethylphosphino)ethane] to form Pd(SeC6H4Cl)Me(dmpe) (5) and Me-SeC6H4Cl. A second equivalent of (ClC6H4Se)2 reacts with 5 to cleave the second Pd-Me bond to give Pd(SeC6H4Cl)2(dmpe) (6) and Me-SeC6H4Cl. Similarly, PdMeTol(dmpe) (7) (Tol=4-tolyl) forms predominantly Pd(SeC6H4Cl)Tol(dmpe) (8) together with some Pd(SeC6H4Cl)Me(dmpe) (5), and 8 reacts with (ClC6H4Se)2 to form Pd(SeC6H4Cl)2(dmpe) (6) and Tol-SeC6H4Cl. Bis(4-chlorophenyl) diselenide reacts with PtTol2(bpy) (9) (Tol=4-tolyl) to form Pt(SeC6H4Cl)2Tol2(bpy) (10) which, together with 2, has a trans-configuration for the selenolate ligands. X-ray structural studies of octahedral 10 as the solvate 10 · 3CHCl3 and square planar 5 are reported.  相似文献   

3.
The replacement of SO42− ions by monovalent ions in mineral SrSO4 crystals was investigated under hydrothermal conditions by using aqueous solutions bearing F and OH ions. Experiments were conducted at various temperatures (150-250 °C) for different reaction intervals (1-96 h), with M/SO42− molar ratios of 1, 5 and 10, where M=F or OH. The celestite crystals were completely converted into SrF2 crystals, at 200 °C using a F/SO42− molar ratio=5 for 24 h. The morphology of the converted SrF2 crystals indicated that the heteroionic conversion proceeded by a pseudomorphic replacement process, because the transformed crystals maintained their original shape and dimensions. In contrast, the SrSO4 crystals were instantaneously converted into the Sr(OH)2 phase by a bulk dissolution-recrystallization mechanism, resulting in the formation of large transparent acicular Sr(OH)2 crystals. The differences on the conversion process are mainly associated with the chemical interaction between the mineral crystal and the hydrothermal fluid. In addition, the chemical stability of the converted phase with low solubility is also essential for the heteroionic conversion to proceed by the pseudomorphic replacement process.  相似文献   

4.
A novel conjugation-elongated bis(ethylenedithio)tetraselenafulvalene (BETS) type donor, 2,5-bis(4,5-ethylenedithio-1,3-diselenol-2-ylidene)-2,3,4,5-tetrahydrothiophene (BEDT-HBDST) and its magnetic and non-magnetic anion salts, (BEDT-HBDST)2MX4 (MX4=FeCl4, GaCl4, FeBr4 and GaBr4), were prepared. These four salts are isostructural and belong to the space group of P2/c. They showed semiconducting behavior with small activation energies (59-64 meV). The band structures of these salts are quasi one-dimensional and there is a midgap between the upper band and the lower band, since the degree of dimerization is significant in the stacking direction. The MX4 ions are located between the donor columns and near to the ethylenedithio moieties of the donor molecules. The magnetic susceptibilities of the FeCl4 and FeBr4 salts follow the Curie-Weiss law with Curie constants of 4.6 and 4.8 emu K mol−1 (sum of the spins of S=5/2 and S=1/2) and negative Weiss temperatures of θ=−1.2 and −4.9 K, respectively, revealing a weak antiferromagnetic interaction of 3d spins of the FeCl4 and FeBr4 anions. The Fe?Fe (6.66-7.60 Å), Cl?Cl (4.81-4.82 Å) and Br?Br (4.74-4.77 Å) distances in the crystal structures of these salts are significantly long. Therefore, the direct magnetic interaction between the 3d spins of the nearest neighboring Fe3+ ions appears to be not readily accessible.  相似文献   

5.
cis-(η5-MeC5H4)W(CO)2P(OiPr)3I (1) was converted to the trans isomer 2 in the solid state (90-110 °C). The reaction was monitored by heating 1 in NMR tubes for periods of time (2-60 min), cooling the tubes to room temperature and determining the conversion by solution 31P and 1H NMR spectroscopy. The data were consistent with a first-order reaction and yielded an activation energy of 59 ± 3 kJ mol−1. Comparative kinetic data were obtained from an in situ analysis of a powder-XRD study of 1. The powder-XRD study was conducted at 80-100 °C (10-60 min), yielding an activation energy of 52 ± 2 kJ mol−1 (first-order reaction). The reaction could not be monitored by single crystal X-ray diffraction as the crystal disintegrated over time on heating. This disintegration process was monitored by optical microscopy and revealed that while the bulk crystal morphology was retained the crystal surface roughened with time. The compounds 1 and 2 were also structurally characterised by X-ray crystallographic techniques.  相似文献   

6.
Manuela Kim 《Talanta》2007,72(3):1054-1058
A simple and sensitive HPLC post-derivatization method with colorimetric detection has been developed for the determination of N-nitroso glyphosate in samples of technical glyphosate. Separation of the analyte was accomplished using an anionic exchange resin (2.50 mm × 4.00 mm i.d., 15 μm particle size, functional group: quaternary ammonium salt) with Na2SO4 0.0075 M (pH 11.5) (flow rate: 1.0 mL min−1) as mobile phase. After separation, the eluate was derivatized with a colorimetric reagent containing sulfanilamide 0.3% (w/v), [N-(1-naphtil)ethilendiamine] 0.03% (w/v) and HCl 4.5 M in a thermostatized bath at 95 °C. Detection was performed at 546 nm. All stages of the analytical procedure were optimized taking into account the concept of analytical minimalism: less operation times and costs; lower sample, reagents and energy consumption and minimal waste. The limit of detection (k = 3) calculated for 10 blank replicates was 0.04 mg L−1 (0.8 mg kg−1) in the solid sample which is lower than the maximum tolerable accepted by the Food and Agriculture Organization of the United Nations.  相似文献   

7.
Hyphenation of thermogravimetric analyzer (TGA) and thermo-Raman spectrophotometer for in situ monitoring of solid-state reaction in oxygen atmosphere forming NiO-Al2O3 catalyst nanoparticles is investigated. In situ thermo-Raman spectra in the range from 200 to 1400 cm−1 were recorded at every degree interval from 25 to 800 °C. Thermo-Raman spectroscopic studies reveal that, although the onset of formation is around 600 °C, the bulk NiAl2O4 forms at temperatures above 800 °C. The X-ray diffraction (XRD) spectra and the scanning electron microscopy (SEM) images of the reaction mixtures were recorded at regular temperature intervals of 100 °C, in the temperature range from 400 to 1000 °C, which could provide information on structural and morphological evolution of NiO-Al2O3. Slow controlled heating of the sample enabled better control over morphology and particle size distribution (∼20-30 nm diameter). The observed results were supported by complementary characterizations using TGA, XRD, SEM, transmission electron microscopy, and energy dispersive X-ray analysis.  相似文献   

8.
Treatment of SbX3 (X = Br, Cl) with DippnacnacLi (Dippnacnac = [{N(C6H3iPr22,6)C(Me)}2CH]) or Mesnacnac (Mesnacnac = [{N(Mes)C(Me)}2CH], Mes = 2,4,6, trimethyl benzene) affords different products that are dependent on the stoichiometry of the reaction and the halide precursor. When DippnacnacLi is reacted with SbBr3, C-H activation of the ligand backbone is observed and an asymmetric, bridged bromide dimer is isolated. In comparison, the reaction of SbCl3 with MesnacnacLi affords monomeric MesnacnacSbCl2. The solid-state structures were determined using X-ray crystallography.  相似文献   

9.
3,4-Di-(2′-hydroxyethoxy)-4′-nitrobenzylidene II was prepared by condensation reaction of 3,4-dihydroxy-4′-nitrobenzylidene I with 1-chloro-2-ethanol. Monomer II was reacted with p-phenylene diisocyanate to yield polyurethane containing the non-linear optical chromophore 3,4-di-(2′-hydroxyethoxy)-4′-nitrobenzylidene. Polymer III shows thermal stability up to 300 °C in TGA thermogram. Tg value of the polymer obtained from DSC thermogram was 110 °C. The resulting polyurethane III was soluble in common organic solvents such as acetone, DMF and DMSO. The values of electro optic coefficient d33 and d31 of the poled polymer film were 3.15 × 10 −7 and 1.5 × 10 −7 esu, respectively.  相似文献   

10.
The oxidation of glyoxylic acid (HGl) by MnIVL {L4− = tetra deprotonated 1,8-bis(2-hydroxybenzamido)-3,6-diazaoctane} was investigated in the pH range 1.67-10.18, at 25-45 °C and 0.5 M ionic strength. The reaction exhibited biphasic kinetics with MnIIIL as the reactive intermediate. MnIV was reduced to MnII. The products of oxidation of HGl were identified as formic acid and CO2 in acidic medium, and oxalate in basic medium, consistent with the stoichiometry: −Δ[MnIV]/−Δ[HGl] = 1. In acidic medium, both MnIVL and MnIIIL formed outer-sphere adducts with the neutral HGl {HC(OH)2COOH} molecule, with an association constant Qav of 28 and 70 M−1, respectively. A similar adduct formation was not observed for the glyoxylate mono anion {Gl, CH(OH)2(CO2)} and glyoxylate dianion {Gl2−, CH(OH)(O)CO2}. The rate and activation parameters for the various paths are reported and an outer-sphere electron transfer mechanism is suggested.  相似文献   

11.
Alkyl and dialkylammonium tetrafluoroborate promoted cis-trans isomerization of 1,3,5-trimethyl-1,3,5-triphenylcyclotrisiloxane (1) in DMSO-d6 were studied. The isomerization equilibrium constant K are within the range of 3.74-3.30 from 22 to 47 °C. Thermodynamic parameters of ΔH° and ΔS° for the isomerization were −0.95 kcal/mol and −0.59 cal/mol-K respectively. The isomerization rate is first order in [cis-1] and second order in [RnNH4−nBF4]. Both components of RnNH4−n+ and BF4 are essential for the catalytic cis-trans isomerization. The catalytic strength follows the decreasing order of +H3N(CH2)6NH3+>n-C8H17NH3+>n-C16H33NH3+>Me3CNH3+>PhCH2NH3+>Et2NH2+?Ph2CHNH3+, Et3NH+. Inversion region was observed in the plot of ln(kf/T) versus (1/T) with the ceiling located at around 38 °C. The positive activation enthalpy of 9 kcal/mol was estimated at 22-32 °C. The activation enthalpy turns to be slightly negative at T>38 °C.  相似文献   

12.
The oxidation of a series of substituted pyridines by dimethyldioxirane (1) produced the expected N-oxides in quantitative yields. The second order rate constants (k2) for the oxidation of a series of substituted pyridines (2a-g) by dimethyldioxirane were determined in dried acetone at 23 °C. An excellent correlation with Hammett sigma values was found (ρ = −2.91, r = 0.995). Kinetic studies for the oxidation of 4-trifluoromethylpyridine by 1 were carried out in the following dried solvent systems: acetone (k2 = 0.017 M−1 s−1), carbon tetrachloride/acetone (7:3; k2 = 0.014 M−1 s−1), acetonitrile/acetone (7:3; k2 = 0.047 M−1 s−1), and methanol/acetone (7:3; k2 = 0.68 M−1 s−1). Kinetic studies of the oxidation of pyridine by 1 versus mole fraction of water in acetone [k2 = 0.78 M−1 s−1 (χ = 0) to k2 = 11.1 M−1 s−1 (χ = 0.52)] were carried out. The results showed the reaction to be very sensitive to protic, polar solvents.  相似文献   

13.
The pinacol coupling of benzaldehyde (0.25 M or 1.25 M) in water was catalyzed by 5-25 mol % CrCl2 in the presence of Zn-dust or Al-dust at 20 °C or 60 °C. In all cases at most 50% of the pinacol coupling product, 1,2-diphenyl-1,2-ethanediol, was obtained with the major product, benzyl alcohol, being formed by a competitive 2e reduction of the carbonyl. The dl- to meso-diastereoselectivity of the pinacol products ranged from 0.6:1 to nearly 1:1.  相似文献   

14.
A detailed analysis of the 35Cl/37Cl isotope effects observed in the 19.11 MHz 103Rh NMR resonances of [RhCln(H2O)6−n]3−n complexes (n = 3–6) in acidic solution at 292.1 K, shows that the ‘fine structure’ of each 103Rh resonance can be understood in terms of the unique isotopologue and in certain instances the isotopomer distribution in each complex. These 35Cl/37Cl isotope effects in the 103Rh NMR resonance of the [Rh35/37Cl6]3− species manifest only as a result of the statistically expected 35Cl/37Cl isotopologues, whereas for the aquated species such as for example [Rh35/37Cl5(H2O)]2−, cis-[Rh35/37Cl4(H2O)2] as well as the mer-[Rh35/37Cl3(H2O)3] complexes, additional fine-structure due to the various possible isotopomers within each class of isotopologues, is visible. Of interest is the possibility of the direct identification of stereoisomers cis-[RhCl4(H2O)2], trans-[RhCl4(H2O)2], fac-[RhCl3(H2O)3] and mer-[RhCl3(H2O)3] based on the 103Rh NMR line shape, other than on the basis of their very similar δ(103Rh) chemical shift. The 103Rh NMR resonance structure thus serves as a novel and unique ‘NMR-fingerprint’ leading to the unambiguous assignment of [RhCln(H2O)6−n]3−n complexes (n = 3–6), without reliance on accurate δ(103Rh) chemical shifts.  相似文献   

15.
A gas chromatography–tandem mass spectrometric (GC–MS/MS) method has been established for the determination of cyanide in surface water. This method is based on the derivatization of cyanide with 2-(dimethylamino)ethanethiol in surface water. The following optimum reaction conditions were established: reagent dosage, 0.7 g L−1 of 2-(dimethylamino)ethanethiol; pH 6; reaction carried out for 20 min at 60 °C. The organic derivative was extracted with 3 mL of ethyl acetate, and then measured by using GC–MS/MS. Under the established conditions, the detection and quantification limits were 0.02 μg L−1 and 0.07 μg L−1 in 10-mL of surface water, respectively. The calibration curve had a linear relationship relationship with y = 0.7140x + 0.1997 and r2 = 0.9963 (for a working range of 0.07–10 μg L−1) and the accuracy was in a range of 98–102%; the precision of the assay was less than 7% in surface water. The common ions Cl, F, Br, NO3, SO42−, PO43−, K+, Na+, NH4+, Ca2+, Mg2+, Ba2+, Mn4+, Mn2+, Fe3+, Fe2+ and sea water did not interfere in cyanide detection, even when present in 1000-fold excess over the species. Cyanide was detected in a concentration range of 0.07–0.11 μg L−1 in 6 of 10 surface water samples.  相似文献   

16.
Four copper(II) complexes were synthesized by reactions of new imidazole-containing polyamine ligand N1-(2-aminoethyl)-N1-(1H-imidazol-4-ylmethyl)-ethane-1,2-diamine (HL) with Cu(ClO4)2 · 6H2O under different pH and their structures were characterized by X-ray crystallography. Interestingly, the complexes have diverse structures from protonated ligand [H3(HL)][CuCl4] · Cl (1), dinuclear [Cu2(HL)2Cl](ClO4)3 · H2O (2), one-dimensional chain polynuclear {[Cu(L)](ClO4)}n (3) to cyclic-tetranuclear [Cu4(L)4](ClO4)4 · 3CH3CN (4) coordination compounds by varying reaction pH from acidic to basic. The results indicate that the reaction pH has great impact on the formation and structure of the complexes. The magnetic measurements show that there are antiferromagnetic interactions between the Cu(II) centers with g = 2.09, J = −39.0 cm−1 and g = 2.17, J = −36.8 cm−1 for 3 and 4, respectively.  相似文献   

17.
Dynamic 1H NMR (500 MHz) investigation of 4-methylphenoxyimidoyl azides (4-CH3-C6H4-O-CN-Y)-N3, Y=4-CH3-C6H4-SO2-, 4-Br-C6H4-SO2-, C6H5SO2-, CH3-SO2-, -CN in acetone-d6 at temperature range of 195-280 K is reported. The observed free energy barrier (almost 12 kcal mol−1) is attributed to conformational isomerisation about the N-S bond for Y=4-CH3-C6H4-SO2-, 4-Br-C6H4-SO2-, C6H5SO2-, CH3-SO2- and (almost 14 kcal mol−1) to configurational isomerisation (E/Z) about CN bond for Y=-CN.  相似文献   

18.
Exchange of PMe2Ph for PPh3 in (η5-pentadienyl)ruthenium{bis(triphenylphosphine)}chloride, (η5-C5H7)Ru(PPh3)2Cl (1) under first order conditions proceeds rapidly in THF at room temperature. A pseudo-first order rate constant of 17 ± 2 × 10−4 s−1 is obtained for the reaction at 21 °C. The rate constant is essentially independent of the phosphine concentration. The activation parameters, ΔH = 16.1 ± 0.4 kcal mol−1 and ΔS = −16 ± 1 cal K−1 mol−1 differ from those reported for phosphine exchange in CpRu(PPh3)2Cl (2) and (η5-indenyl)Ru(PPh3)2Cl (3). The reaction of 1 with PMe2Ph is about 70 times faster than the reaction of 2 at 30 °C and some 40 times faster than the reaction of 3 at 20 °C. (η5-C5H7)Ru(PPh3)2Cl(1) is more active than the ruthenium(II) complexes 2, 3, and TpRu(PPh3)2Cl (4) in the catalytic dimerization of terminal alkynes with nearly quantitative conversion of PhCCH and FcCCH at ambient temperature in 24 h. The enhanced substitution rate is accompanied by >50% conversion of phenylacetylene to oligomeric products. Reaction of 1 with NaPF6 in acetonitrile yields the cationic ruthenium(II) complex [(η5-C5H7)Ru(PPh3)2(CH3CN)][PF6] (7). The latter complex is much less active in reactions with phenylacetylene than 1 but avoids the formation of oligomeric products.  相似文献   

19.
A coordination polymer [Cu(nip)(phen)]n was hydrothermally synthesized by the reaction of Cu(NO3)2 with 5-nitroisophthalic acid and phen. Single-crystal structure analysis showed that the complex crystallized in the monoclinic space group P21/c; a = 10.6566(13); b = 12.5931(15); c = 13.0514(16) Å; β = 95.474(2)°, V = 1743.5(4) Å3; Z = 4. The standard molar enthalpy of formation of the complex was determined to be −554 ± 11 kJ mol−1.  相似文献   

20.
Platinum complexes of the type [Pt(cis-1,4-DACH)(L)2]X, where cis-1,4-DACH = cis-1,4-diaminocyclohexane; L = adenine (ade) (1), hypoxanthine (hyp) (2), 9-methylguanine (9-megua) (3), cytosine (cyt) (4), or 1-methylcytosine (1-mecyt) (5); and X = SO4 or Cl2 groups, were synthesized and characterized by elemental analysis and by 1H, 13C, and 195Pt nuclear magnetic resonance spectroscopy. The crystals of [Pt(cis-1,4-DACH)(9-megua)2]SO4[9-megua-H]2SO4 (3) and [Pt(cis-1,4-DACH)(1-mecyt)2]Cl2 · 6H2O (5) were also subjected to single-crystal X-ray diffraction. The base/PtN4 coordination plane dihedral angles were 74.55° and 85.61° in complex 3 and 78.12° and 81.80° in complex 5. The platinum had distorted square planar geometry in both complexes; the two adjacent corners were occupied by the two nitrogen atoms of cis-1,4-DACH, and the other two corners were occupied by the two N7 atoms of 9-megua in complex 3 and the two N3 atoms of 1-mecyt in complex 5. The cis-1,4-DACH, which has a unique twist-boat configuration, formed a seven-member chelating ring with platinum, which led to considerable strain during bidentate cis-1,4-DACH binding. Cations of both complexes 3 and 5 adopted C2 molecular symmetry. These adducts were the models for the intrastand cross-links that were relevant to the binding of the Pt(II) antitumor drugs to DNA.  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