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.
Solubility of CO2 in six hydroxyl ammonium ionic liquids 2-hydroxyethanaminium acetate [hea], bis(2-hydroxyethyl)ammonium acetate [bheaa], 2-hydroxy-N-(2-hydroxyethyl)-N-methylethanaminium acetate [hhemea], 2-hydroxyethanaminium lactate [hel], bis(2-hydroxyethyl)ammonium lactate [bheal], 2-hydroxy-N-(2-hydroxyethyl)-N-methylethanaminium lactate [hhemel] at temperatures (298.15, 313.15, and 328.16) K and pressures ranging from (100 to 1600) kPa was determined. From the experimental solubility data, the Henry’s constant of CO2 for each hydroxyl ammonium ionic liquids was estimated and reported as a function of temperature. Furthermore, enthalpy and entropy of absorption were obtained from estimated Henry’s constant. The results showed that the solubility increase with increasing pressure and decrease with increasing temperature and the solubility of CO2 in these six hydroxyl ammonium ionic liquids was in sequence: [hea] > [bheaa] > [hel] > [bheal] > [hhemel] > [hhemea].  相似文献   

2.
The effect of cationic micelles of cetyltrimethylammonium bromide (CTAB) on the interaction of dipeptide glycyl-tyrosine (Gly-Tyr) with ninhydrin under varying conditions has been studied spectrophotometrically at 70 °C and pH 5.0. The reaction followed first- and fractional-order kinetics with respect to [Gly-Tyr] and [ninhydrin], respectively. Increase in total concentration of CTAB from 0 to 70 × 10−3 mol dm−3 resulted in an increase in the pseudo-first-order rate constant (kψ) by a factor of ca. 3. Quantitative kinetic analysis of kψ  [CTAB] data was performed on the basis of pseudo-phase model of the micelles (proposed by Menger and Portnoy and developed by Bunton) and Piszkiewicz model. A possible mechanism has been proposed and the kinetic data have been used to evaluate the micellar binding constants KS (268 mol−1 dm3 for Gly-Tyr) and KN (64 mol−1 dm3 for ninhydrin).  相似文献   

3.
The kinetics and mechanism of Ag(I)-catalyzed oxidation of l-alanine by cerium (IV) in sulfuric acid media have been investigated by titrimetric technique of redox in the temperature range of 298–313 K. It is found that the reaction is of first order with respect to Ce(IV) and l-alanine, and it is of a positive fractional order with respect to Ag(I). It is found that the pseudo first order ([l-alanine] ? [Ce(IV)] ? [Ag(I)]) rate constant k′ increases with the increase of[H+]. The major oxidation product of alanine has been identified as acetaldehyde by an 1H NMR and IR spectroscopy. Under the experimental conditions, the kinetically active species has been found to be Ce4+. Under nitrogen atmosphere, the reaction system can initiate the polymerization of acrylonitrile, indicating generation of free radicals. On the basis of the experimental results, a suitable mechanism has been proposed. The rate constants of the rate-determining step together with the activation parameters were evaluated.  相似文献   

4.
The kinetics of the osmium(VIII) (Os(VIII)) catalyzed oxidation of diclofenac sodium (DFS) by diperiodatocuprate(III) (DPC) in aqueous alkaline medium has been studied spectrophotometrically at a constant ionic strength of 1.0 mol⋅dm−3. The reaction showed first order kinetics in [Os(VIII)] and [DPC] and less than unit order with respect to [DFS] and [alkali]. The rate decreased with increase in [periodate]. The reaction between DFS and DPC in alkaline medium exhibits 1:2 [DFS]:[DPC] stoichiometry. However, the order in [DFS] and [OH] changes from first order to zero order as their concentration increases. Changes in the ionic strength and dielectric constant did not affect the rate of reaction. The oxidation products were identified by LC-ESI-MS, NMR, and IR spectroscopic studies. A possible mechanism is proposed. The reaction constants involved in the different steps of the mechanism were calculated. The catalytic constant (K C) was also calculated for Os(VIII) catalysis at the studied temperatures. From plots of log 10 K C versus 1/T, values of activation parameters have been evaluated with respect to the catalytic reaction. The activation parameters with respect to the slow step of the mechanism were computed and discussed, and thermodynamic quantities were also determined. The active osmium(VIII) and copper(III) periodate species have been identified.  相似文献   

5.
《Tetrahedron: Asymmetry》2006,17(14):2069-2073
Four commercially available lipases and two native lipases from Aspergillus niger AC-54 and Aspergillus terreus AC-430 were used for the resolution of (RS)-Ibuprofen in systems containing the ionic liquids [BMIM][PF6] and [BMIM][BF4]. The lipases showed higher conversion in a two-phase system using [BMIM][PF6] and isooctane compared to that in pure isooctane. Although the best enzyme was a commercially available lipase from Candida rugosa (E = 8.5), another native lipase, produced in our laboratory, from A. niger gave better enantioselectivity (E = 4.6) than the other lipases tested (E = 1.9–3.3.). After thorough optimization of several reaction conditions (type and ratios of isooctane/ionic liquid, amount of enzyme, and reaction time), the E-value of A. niger lipase (15% w/v) could be duplicated (E = 9.2) in a solvent system composed of [BMIM][PF6] and isooctane (1:1) after 96 h of reaction.  相似文献   

6.
The charge accumulation due to peroxidase (POD)-catalyzed reduction of H2O2 in a test solution (4 μL) by Os(II) in a POD/PVI[Os(dmebpy)2Cl]-immobilized layer on an electrode (PVI = poly(1-vinylimidazole), dmebpy = 4,4′-dimethyl-2,2′-bipyridine) was monitored potentiometrically for the detection of H2O2. Before potentiometry, the Os(II)/Os(III) ratio of the modified electrode was controlled by pre-electrolysis at a given potential in a separated electrolysis cell. The redox potential of the Os polymer film in the test solution shifted to the positive side on the addition of H2O2 and reached a constant value due to the accumulation of Os(III) in the film. The total amount of the accumulated charge was determined from the area of the portion corresponding to the redox potential shift on a reversible cyclic voltammogram recorded separately. The low detection limit (5 pmol H2O2) was realized with 82–90% of the recovery percentage.  相似文献   

7.
We observed the Raman spectra of carriers, positive polarons and bipolarons, generated in a poly(2,5-bis(3-tetradecylthiophen-2-yl)thieno[3,2-b]thiophene) (PBTTT-C14) film by FeCl3 vapor doping. Electrical conductivity and Raman measurements indicate that the dominant carriers in the conducting state were bipolarons. We identified positive polarons and bipolarons generated in an ionic-liquid-gated transistor (ILGT) fabricated with PBTTT-C14 as an active semiconductor and an ionic liquid 1-butyl-3-methylimidazolium bis(trifluoromethylsulfonyl) imide [BMIM][TFSI] as a gate dielectric using Raman spectroscopy. The relationship between the source−drain current (ID) at a constant source−drain voltage (VD) and the gate voltage (VG) was measured. ID increased above −VG = 1.1 V and showed a maximum at −VG = 2.0 V. Positive polarons were formed at the initial stage of electrochemical doping (−VG = 0.8 V). As ID increased, positive bipolarons were formed. Above VG = −2.0 V, bipolarons were dominant. The charge density (n), the doping level (x), and the mobility of the bipolarons were calculated from the electrochemical measurements. The highest mobility (μ) of bipolarons was 0.72 cm2 V−1 s−1 at x = 110 mol%/repeating unit (−VG = 2.0 V), whereas the highest μ of polarons was 4.6 × 10−4 cm2 V−1 s−1 at x = 10 mol%.  相似文献   

8.
《Polyhedron》2005,24(16-17):2165-2172
Five new hydrogen-bonded solvated iron(II) complexes of pyrazolyl- and imidazolyl-based N,N-chelating ligands have been synthesised. Water to ligand-NH hydrogen-bonded bridges occur in the pseudo-dimeric complexes {cis-[Fe(pypzH)2(NCX)2]2(μ-OH2)(H2O)2} · H2O · MeOH (where X = S or Se), and in the chain complex {cis-[Fe(pypzH)2(NCS)2](μ-OH2)}n. A “half” spin-crossover (Tc = 125 K) was observed in the dimeric X = Se complex by means of magnetic measurements and no thermal hysteresis occurred between 4 and 300 K. The crystal structure at 123 K showed Fe–N distances consistent with the magnetism. Each Fe in the dimeric unit was structurally equivalent in the HS–LS state. Removal of the solvate molecules led to HS–HS behaviour over the temperature range 4–300 K. The pseudo-dimer with X = S also showed HS–HS behaviour as did the monomeric analogue cis-[Fe(pypzH)2(NCS)2]H2O and a structurally different methanol-bridged dimer {cis-[Fe(pyimH)2(NCS)2]2(μ-MeOH)2} · 2MeOH (pypzH = 2-(1H-pyrazol-3-yl)-pyridine; pyimH = 2-(1H-imidazol-2-yl)-pyridine).  相似文献   

9.
Six Ru(II) complexes of formula [Ru(L)2(PPh3)2] have been prepared where LH = 4-(aryl)thiosemicarbazones of thiophen-2-carbaldehyde. X-ray crystal structures of five of the complexes are reported. In all the complexes ruthenium is six coordinate with a distorted octahedral cis-P2, cis-N2, trans-S2 donor environment, and each of the two thiosemicarbazone ligands are coordinated in a bidentate fashion forming a four membered chelate ring. The complexes undergo a one-electron oxidation at ~0.5 V vs. Ag/AgCl. The EPR spectrum of the electrochemically oxidized solution at 100 K shows a rhombic signal, with transitions at g1 = 2.27, g2 = 2.00 and g3 = 1.80. DFT calculations on one of the complexes suggest that there is 35% ruthenium and 17% sulfur orbital contribution to the HOMO. These results suggest that the assignment of metal atom oxidation states in these compounds is not unambiguous.  相似文献   

10.
《Comptes Rendus Chimie》2014,17(5):477-483
The ligand Hbpq = N-(8-quinolyl)pyridine-2-carboxamide) has been prepared using tetrabutylammonium bromide (TBAB) as an environmentally friendly reaction medium. Four new complexes of this ligand, [M(bpq)X] (M = Cu(II), X = SCN̄ (1), N3̄ (2); M = Ni(II), X = SCN̄ (3), N3̄ (4)), have also been synthesized and fully characterized. The crystal and molecular structures of [Cu(bpq)(NCS)]n (1) have been determined by X-ray crystallography. Copper(II) ion adopts a distorted square pyramidal (4 + 1) coordination in this complex. Hbpq ligand shows a strong emission at 500 nm in acetonitrile solution. The emission is quenched in the presence of copper(II) acetate, apparently because of the formation of [Cu(L)(OAc)(H2O)] complex. Introduction of nitric oxide (NO) into the acetonitrile solution at room temperature induces an increase in the fluorescence intensity, presumably due to the reduction of Cu(II) to Cu(I). This process is reversible and can form a basis for direct detection of NO.  相似文献   

11.
In the present work we report a reaction in which dimethyl sulfoxide, initially used as solvent, undergoes oxidation to form sulfate, which then participates to the formation of a linear one-dimensional copper chain. Indeed, using [Cu(bipy)Cl2], a building block largely applied in synthesis of molecular magnetic compounds, the coordination compound [Cu(bipy)(H2O)2(SO4)]n, where bipy = 2,2′-bipyridine was obtained. Magnetic characterization of complex shows a weak antiferromagnetic interaction between the copper(II) ions with J = ?0.53 cm?1. DFT calculations demonstrate that the pathway for the weak antiferromagnetic interaction is through the sulfate bridge.  相似文献   

12.
《Comptes Rendus Chimie》2015,18(7):758-765
The catalytic peroxidative oxidation (with H2O2) of cyclohexane in an ionic liquid (IL) using the tetracopper(II) complex [(CuL)24-O,O′,O′′,O′′′-CDC)]2·2H2O [HL = 2-(2-pyridylmethyleneamino)benzenesulfonic acid, CDC = cyclohexane-1,4-dicarboxylate] as a catalyst is reported. Significant improvements on the catalytic performance, in terms of product yield (up to 36%), TON (up to 529), reaction time, selectivity towards cyclohexanone and easy recycling (negligible loss in activity after three consecutive runs), are observed using 1-butyl-3-methylimidazolium hexafluorophosphate as the chosen IL instead of a molecular organic solvent including the commonly used acetonitrile. The catalytic behaviors in the IL and in different molecular solvents are discussed.  相似文献   

13.
Dissociative ionization of 1,4-bis(2,5-phenyloxazolyl)benzene (POPOP) molecule by electron impact in gaseous phase is studied. Potentials of appearance of some fragments of the molecule under study are determined from the experimentally measured dependences of ionization cross-section on the ionizing electron energy. A new ion with m/z = 144 [C9H6ON]+ is detected in the mass spectrum of the POPOP molecule, being complementary to the fragment with m/z = 220 [C15H10ON]+. The threshold of appearance of this ion is determined (Eap = 9.51 eV) as well as the first ionization potential of the POPOP molecule and fragment ion appearance potentials.  相似文献   

14.
The kinetics of oxidation of a non-steroidal analgesic drug, aspirin (ASP) by diperiodatocuprate(III)(DPC) in the presence and absence of osmium(VIII) have been investigated at 298 K in alkaline medium at a constant ionic strength of 0.10 mol dm−3 spectrophotometrically. The reaction showed a first-order in [DPC] and less than unit order in [ASP] and [alkali] for both the osmium(VIII) catalysed and uncatalysed reactions. The order with respect to Os(VIII) concentration was unity. The effects of added products, ionic strength, periodate and dielectric constant have been studied. The stoichiometry of the reaction was found to be 1:4 (ASP:DPC) for both the cases. The main oxidation product of aspirin was identified by spot test, IR, NMR and GC–MS. The reaction constants involved in the different steps of the mechanisms were calculated for both reactions. Activation parameters with respect to slow step of the mechanisms were computed and discussed for both the cases. The thermodynamic quantities were also determined for both reactions. The catalytic constant (KC) was also calculated for catalysed reaction at different temperatures and the corresponding activation parameters were determined.  相似文献   

15.
The kinetics of the diazotization reaction of procaine in the presence of anionic micelles of sodium dodecyl sulfate (SDS) and cationic micelles of cetyltrimethyl ammonium bromide (CTAB), dodecyltrimethyl ammonium bromide (DDTAB) and tetradecyltrimethyl ammonium bromide (TDTAB) were carried out spectrophotometrically at λmax = 289 nm. The values of the pseudo first order rate constant were found to be linearly dependent upon the [NaNO2] in the concentration range of 1.0 × 10−3 mol dm−3 to 12.0 × 10−3 mol dm−3 in the presence of 2.0 × 10−2 mol dm−3 acetic acid. The concentration of procaine was kept constant at 6.50 × 10−5 mol dm−3. The addition of the cationic surfactants increased the reaction rate and gave plateau like curve. The addition of SDS micelles to the reactants initially increased the rate of reaction and gave maximum like curve. The maximum value of the rate constant was found to be 9.44 × 10−3 s−1 at 2.00 × 10−3 mol dm−3 SDS concentration. The azo coupling of diazonium ion with β-naphthol (at λmax = 488) nm was found to linearly dependent upon [ProcN2+] in the presence of both the cationic micelles (CTAB, DDTAB and TDTAB) and anionic micelles (SDS). Both the cationic and anionic micelles inhibited the rate of reactions. The kinetic results in the presence of micelles are explained using the Berezin pseudophase model. This model was also used to determine the kinetic parameters e.g. km, Ks from the observed results of the variation of rate constant at different [surfactants].  相似文献   

16.
《Polyhedron》2005,24(16-17):2491-2496
The bis(dithiolene) oxovanadium complex, namely (n-Bu4N)2[(dcbdt)2VO] (dcbdt = 4,5-dicyanobenzene-1,2-dithiolato), was unprecedentedly obtained from the reaction of Na2dcbdt and vanadium trichloride. An X-ray structure analysis indicated that [(dcbdt)2VO] moieties are surrounded by n-Bu4N cations and there was no direct interaction among these moieties. Although there was no direct interaction among them, several S⋯H, C⋯H and N⋯H van der Waals contacts between n-Bu4N cations and [(dcbdt)2VO] moieties were observed. The ESR and SQUID measurement showed that the vanadium atom is in the state of V(IV) (S = 1/2) and these magnetic moments interact with each other very weakly antiferromagnetically (θ = -0.371 K).  相似文献   

17.
A new amino acid ionic liquid (AAIL) [C3mim][Val] (1-propyl-3-methylimidazolium valine) was prepared by the neutralization method. Using the solution-reaction isoperibol calorimeter, molar solution enthalpies of the ionic liquid [C3mim][Val] with known amounts of water and with different concentrations in molality were measured at T = 298.15 K. In terms of standard addition method (SAM) and Archer’s method, the standard molar enthalpy of solution for [C3mim][Val] without water, ΔsHm = (−55.7 ± 0.4) kJ · mol−1, was obtained. The hydration enthalpy of the cation [C3mim]+, ΔH+ ([C3mim]+) = −226 kJ · mol−1, was estimated in terms of Glasser’s theory. Using the RD496-III heat conduction microcalorimeter, the molar enthalpies of dilution, ΔDHm(mi  mf), of aqueous [C3mim][Val] with various values of molality were measured. The values of ΔDHm(mi  mf) were fitted to Pitzer’s ion-interaction model and the values of apparent relative molar enthalpy, φL, calculated using Pitzer’s ion-interaction model.  相似文献   

18.
The kinetics of Ru(III) catalysed oxidation of l-leucine by diperiodatoargentate(III) (DPA) in alkaline medium at 298 K and a constant ionic strength of 0.60 mol dm−3 was studied spectrophotometrically. The oxidation products are pentanoic acid and Ag(I). The stoichiometry is [l-leucine]:[DPA] = 1:2. The reaction is of first order in Ru(III) and [DPA] and has less than unit order in both [l-leu] and [alkali]. The oxidation reaction in alkaline medium has been shown to proceed via a Ru(III)–l-leucine complex, which further reacts with one molecule of monoperiodatoargentate(III) (MPA) in a rate determining step followed by other fast steps to give the products. The main products were identified by spot test and spectral studies. The reaction constants involved in the different steps of the mechanism are calculated. The catalytic constant (Kc) was also calculated for the Ru(III) catalysed reaction at different temperatures. From the plots of log Kc versus 1/T, values of activation parameters with respect to the catalyst have been evaluated. The activation parameters with respect to the slow step of the mechanism are computed and discussed, and thermodynamic quantities are also determined. The active species of catalyst and oxidant have been identified.  相似文献   

19.
Reactions of Schiff bases (H2apahR) derived from acetophenone and acid hydrazides, triethylamine and [Ru(PPh3)3Cl2] (1:2:1 mole ratio) in methanol provide cyclometallated ruthenium(III) complexes of formula trans-[Ru(apahR)(PPh3)2Cl] in 74–81% yields. The complexes have been characterized by elemental analysis, magnetic susceptibility, spectroscopic (infrared, electronic and EPR) and electrochemical measurements. X-ray crystal structures of two representative complexes have been determined. In each complex, the metal centre is in distorted octahedral CNOClP2 coordination sphere assembled by the C,N,O-donor meridionally spanning apahR2?, the chloride and the two mutually trans-oriented PPh3 molecules. All the complexes are one-electron paramagnetic (μeff. = 1.85–1.98 μB) and display rhombic EPR spectra in frozen (120 K) dichloromethane-toluene (1:1) solution. Electronic spectra of the complexes display several absorptions within 470–270 nm due to ligand-to-metal charge transfer and ligand centred transitions. The complexes are redox active and display a Ru(III)  Ru(II) reduction and a Ru(III)  Ru(IV) oxidation in the potential ranges ?0.66 to ?0.70 V and 0.75 to 0.80 V (vs. Ag/AgCl), respectively.  相似文献   

20.
A nonmagnetic compound, [NO2BzPy][Cu(mnt)2] (mnt2? = maleonitriledithiolate; NO2BzPy+ = 1-(4′-nitrobenzyl)pyridinium), is isostructural with [NO2BzPy][Ni(mnt)2], which is a quasi-1D spin system and exhibits a spin-Peierls-like transition with J = 192 K in the gapless state and spin energy gap = 738 K in the dimerization state, respectively. Further, five nonmagnetic impurity doped compounds [NO2BzPy][CuxNi1?x(mnt)2] (x = 0.04–0.74) were prepared, and their crystal structures as well as magnetic properties were investigated. The nonmagnetic doping causes the suppression of the spin transition with an average rate of 139(13) K/percentage of dopant concentration, and the transition collapse is estimated at around x > 0.5.  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