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.
Gas electron diffraction is applied to determine the geometric parameters of the silacyclobutane molecule using a dynamic model where the ring puckering was treated as a large amplitude motion. The structural parameters and the parameters of the potential function were refined taking into account the relaxation of the molecular geometry estimated from ab initio calculations at the MP2/6-311+G(d, p) level of theory. The potential function has been described as V() = V0[(/e)2 − 1]2 with the following parameters V0 = 0.82 ± 0.60 kcal/mol and e = 33.5 ± 2.7°, where is a puckering angle of the ring.

The geometric parameters at the minimum V() (ra in Å, in degrees and uncertainties given as three times the standard deviations including a scale error) are: r(Si–Hax) = 1.467(96), r(Si–Heq) = 1.468(96), r(Si–C) = 1.885(2), r(C–C) = 1.571(3), r(C–H) = 1.100(3), CSiC = 77.2(9), HSiH = 108.3, SiCHeq = 123.5(16), SiCHax = 111.9(16), CC5Heq = 118.4(24), CC5Hax = 112.3(24), HC3H = 107.7, δ(HSiH) = 6.6, δ(HC3H) = 7.0, where the tilts δ, HSiH, and HC3H are estimated from ab initio constraints. The structural parameters are compared with those obtained for related compounds.  相似文献   


2.
Fan J  Wang J  Ye C 《Talanta》1998,46(6):1285-1292
The acid dissociation constants (Ka), base dissociation constants (Kb) and the autoprotolysis constants (Ks) for 2,2′-bipyridyl in water and in water+alcohol(methanol, ethanol, iso-propanol) mixed solvents have been determined at 25°C and an ionic strength of 0.1 mol l−1, from a direct potentiometric method based on the treatment of the data of a single pH titration. It has been shown that Ka increases, whereas Kb and Ks decrease, with increasing proportion of the alcohol in the mixed solvents. Linear relations between pKa, pKb, pKs and the mole fraction of the alcohol were observed in the composition range investigated. These results are discussed in terms of the properties of solvent and the interactions of the different species existing in dissociation equilibrium with solvents. It is concluded that the higher stabilization of both 2,2′-bipyridyl and its protonated form by dispersion forces and of the proton by its interaction with solvent molecules in the mixed solvents compared with that in water are largely responsible for the observed changes of pKa with composition. On the other hand, the low stabilization of OH in the mixed solvents relative to that in water and the electrostatic effect are the main factors in determining the solvent effect on pKb.  相似文献   

3.
Three interpenetrated polymeric networks, {[Co(bpp)(OH-BDC)] · H2O}n (1) [Ni(bpp)1.5(H2O)(OH-BDC)]n (2) and {[Cd(bpp)(H2O)(OH-BDC)] · 2H2O}n (3), have been prepared by hydrothermal reactions of 1,3-bis(4-pyridyl)propane (bpp), 5-hydroxyisophthalic acid (OH-H2BDC), with Co(NO3)2 · 6H2O, Ni(NO3)2 · 6H2O and Cd(NO3)2 · 4H2O, respectively. Single-crystal X-ray diffraction analyses reveal that the three compounds all exhibit interpenetrated but entirely different structures. Compound 1 is a fourfold interpenetrated adamantanoid structure with water molecules as space fillers, in which bpp adopts a TG conformation (T = trans, G = gauche). Compound 2 is an interdigitated structure from the interpenetrated long arms of one-dimensional molecular ladders, while bpp in 2 adopts both TT and TG conformations. Compound 3 is a twofold interpenetrated three-dimensional network from a one-dimensional metal-carboxylate chain bridged by TG conformational bpp. The hydrogen bonding interactions in 1–3 further stabilize the whole structural frameworks and play critical roles in their constructions.  相似文献   

4.
With the aim of understanding the influence of donor solvents on the reactivity of the amine complexes [RuCl2(PPh3)2(piperidine)] (1) and [RuCl2(PPh3)2(imidazole)2] (2) in the presence of ethyldiazoacetate, and on the properties of the resulting polymer, a ring opening metathesis polymerization of norbornene was carried out in the presence of small amounts of common solvents such as additives (isopropanol, THF, N,N-dimethylformamide, 2,6-lutidine, isopropanethiol, acetonitrile, dimethyl sulfoxide, NEt3, NH2Me and pyridine). From observations, typical coordinating solvents like DMSO, NEt3, NH2Me and pyridine, hardly affected the yields when either complex was employed. With other additives, the major advantage was the decrease in the polydispersity indices. On using complex 1 with 2,6-lutidine, observed values of Mw/Mn were as low as 1.3, while the yield decreased from 99% to about 20–30% at RT for 1 min in pure solution. In the case of complex 2, which is almost inactive to ROMP (19% at 50 °C for 5 min with Mw/Mn = 6.30), the yield was three-fold (60% at 50 °C for 5 min with Mw/Mn = 1.95) compared to that of without THF. Further, the Mw/Mn was observed to decrease to 1.34 with 200 eq. of THF.  相似文献   

5.
Ab initio molecular orbital calculations and statistical Monte Carlo simulations employing a combined quantum and molecular mechanical potential were used to determine the enol contents of acetic acid and the acetate ion in aqueous solution. A pKE of 19.3 ± 0.3 was predicted for the keto-enol equilibrium of acetic acid, and 21.8 ± 0.8 for the acetate ion in water. The results are found to be in good accord with Guthrie's calculations based on disproportionation reactions and kinetic data. Combining with the experimental pKa value of acetic acid, we obtained pKak = 26.6 for ionization of acetic acid as a carbon acid in water, and pKaE = 7.3 for ionization of the enol of acetic acid.  相似文献   

6.
Kudo Y  Usami J  Katsuta S  Takeda Y 《Talanta》2003,59(6):1213-1218
Ion-pair formation constants (KMLA mol−1 dm3) of Na+– and K+–18-crown-6 ether (18C6) complexes with MnO4 in water (w) were determined potentiometrically at 25 °C. Simultaneously, extraction constants (Kex mol−2 dm6) of the permanganates with 18C6 from w into 1,2-dichloroethane at 25 °C were obtained from the spectrophotometric determination of distribution ratios of the permanganates. These Kex values were divided into KMLA and other three component equilibrium constants and thereby extraction-selectivity and -ability were discussed in comparison with corresponding metal picrate–18C6 extraction systems reported before.  相似文献   

7.
The mononuclear [Mn(indH)Cl2](CH3OH) (indH: 1,3-bis(2′-pyridylimino)-isoindoline) complex has been prepared and characterized by various techniques such as elemental analysis, IR, UV–vis, ESR spectroscopy and X-ray diffraction. The title compound in the presence of a base such as 1-methylimidazole, imidazole or pyridine is efficient catalyst for the disproportionation of H2O2 in CH3CN. Among the various nitrogenous bases investigated in this study imidazole and substituted imidazoles with strong π-donating ability show better co-catalytic effect.

In case of aqueous solution the complex [Mn(indH)Cl2](CH3OH) shows much higher catalytic activity, and the initial rate of the disproportionation of H2O2 increases with increasing pH and goes through a maximum, which was found at pH  9.6. In this pH value the reaction shows first-order dependence on the catalyst, and saturation kinetics on [H2O2] with Vmax = 8.1 × 10−3 Ms−1, KM = 489 mM, kcat = 38 ± 2 s−1 and k2(kcat/KM) = 79 ± 4 M−1s−1.  相似文献   


8.
The molecular structure of trichloronitromethane has been studied in the gas phase using electron diffraction data. The molecules are found to undergo low barrier rotation about the CN bond with a planar CNO2 moiety in agreement with HF/MP2/B3LYP/6-311G(d,p) calculations. The experimental data are consistent with a dynamic model using a potential function for the torsion of V = (V6/2)(1 − cos 6τ). The major geometrical parameters (rg and ) for the eclipsed form, obtained from least squares analysis of the data are as follows: r(NO3) = r(NO4) = 1.213(2) Å, r(CN) = 1.592(6) Å, r(CCl)av = 1.749(1) Å, Cl5CN/Cl6CN = 109. 6°/106.3°(2), O3NC/O4NC = 117. 6°/114.1°(4), τCl5C1N2O3 = 0.0°, and V6 = 0.20(25) kcal/mol.  相似文献   

9.
The reactions of Zn(NO3)2 · 6H2O and FeSO4 · 7H2O with 4-PDS (4-PDS = 4,4′-dipyridyldisulfide) and NH4SCN in CH3OH afforded the complexes [Zn(NCS)2(4-PDS)]n (1) and [Fe(NCS)2(4-PDS)2 · 4H2O]n (2), respectively, while the reaction of CoCl2 · 6H2O with 4-PDS in CH3OH gave the complex {[Co(4-PDS)2][Cl]2 · 2CH3OH}n, (3). These complexes have been characterized by spectroscopic methods and their structures determined by X-ray crystallography. The 4-PDS ligands in 1 are coordinated to the metal centers through the nitrogen atoms to form 1-D zigzag-chains, and the distorted tetrahedral coordination geometry at each zinc center is completed by a pair of N-bonded thiocyanate ligands. Compound 2 has a 1-D channel-chain structure and each octahedral Fe(II) metal center is coordinated by four 4-PDS ligands and two trans N-bonded thiocyanate ligands. Weak SS interactions in complex 1 link the 1-D chains into 2-D molecular sheets. In complex 2, the channel chains are interlinked through SS interactions to form molecular sheets, which interpenetrate through the SS interactions to form 3-D structures with large cavities that are occupied by the water molecules. Compound 3 also has a 1-D channel-chain structure with each square-planar Co(II) metal center coordinated by four 4-PDS ligands. Multiple C–HCl hydrogen bonds and SO interactions in 3 link the 1-D chains into 2-D structures.  相似文献   

10.
The activity of enzyme I (EI), the first protein in the bacterial PEP:sugar phosphotransferase system, is regulated by a monomer–dimer equilibrium where a Mg2+-dependent autophosphorylation by PEP requires the homodimer. Using inactive EI(H189A), in which alanine is substituted for the active-site His189, substrate binding effects can be separated from those of phosphorylation. Whereas 1 mM PEP (with 2 mM Mg2+) strongly promotes dimerization of EI(H189A) at pH 7.5 and 20 °C, 5 mM pyruvate (with 2 mM Mg2+) has the opposite effect. A correlation between the coupling of N- and C-terminal domain unfolding, measured by differential scanning calorimetry, and the dimerization constant for EI, determined by sedimentation equilibrium, is observed. That is, when the coupling between N- and C-terminal domain unfolding produced by 0.2 or 1.0 mM PEP and 2 mM Mg2+ is inhibited by 5 mM pyruvate, the dimerization constant for EI(H189A) decreases from >108 to <5 × 105 or 3 × 107 M−1, respectively. With 2 mM Mg2+ at 15–25 °C and pH 7.5, PEP has been found to bind to one site/monomer of EI(H189A) with KA′106 M−1G′=−33.7±0.2 kJ mol−1 and ΔH=+16.3 kJ mol−1 at 20 °C with ΔCp=−1.4 kJ K−1 mol−1). The binding of PEP to EI(H189A) is synergistic with that of Mg2+. Thus, physiological concentrations of PEP and Mg2+ increase, whereas pyruvate and Mg2+ decrease the amount of dimeric, active, dephospho-enzyme I.  相似文献   

11.
Binding of Cd2+ by PAMAM 4.5 dendrimer was studied by equilibrium dialysis, isothermal titration calorimetry and zeta-potential measurement. The following binding parameters were obtained: n = 23.8 ± 9.5, Kb = 4.7 ± 0.9 × 103 in water; and n = 41.3 ± 13.4, Kb = 2.1 ± 0.8 × 103 in 0.15 mol/l phosphate-buffered saline. The location of the bound Cd2+ is discussed. The interactions between bovine serum albumin, PAMAM 4.5 dendrimer and cadmium were analyzed using fluorescence and equilibrium dialysis. The competition between Cd2+ binding to BSA and PAMAM 4.5 dendrimer was investigated. It is proposed that PAMAM 4.5 dendrimer could be successfully used for extracting Cd2+ from aqueous solutions (environmental protection).  相似文献   

12.
To study the effect of W concentration and activation temperature of the catalysts a series of WOx/ZrO2 samples with varying concentration of W (10–25 wt.%) were prepared and activated at 650/750 °C. XRD of sample shows 15 wt.% W stabilizes the tetragonal phase of zirconia up to 750 °C. Above and less than 15 wt.% shows peaks corresponding to monoclinic WO3 and monoclinic ZrO2, respectively. Further, the tungsten modification stabilizes the specific surface area of ZrO2. There is an increase in the surface area observed up to 15 wt.% W, which declines on further increase in the concentration. The NH3 TPD confirms the presence of acid sites with varying strength from the broad desorption profile. The 15 wt.% W and activated at 750 °C shows maximum acidity. The results of the nitration reaction of chlorobezene imply the 15 wt.% W and activation at 750 °C shows maximum activity. Not only yield, a better para-selectivity is also achieved with WOx/ZrO2 samples. Effect of activation temperature, W concentration and reaction parameters such as reaction temperature, reaction time, the presence of solvent and solvent free medium on activity and selectivity are studied in details.  相似文献   

13.
Cationic rhodium and iridium complexes of the type [M(COD)(PPh3)2]PF6 (M = Rh, 1a; Ir, 1b) are efficient precatalysts for the hydroformylation of 1-hexene to its corresponding aldehydes (heptanal and 2-methylhexanal), under mild pressures (2–5 bar) and temperatures (60 °C for Rh and 100 °C for Ir) in toluene solution; the linear to branched ratio (l/b) of the aldehydes in the hydroformylation reaction varies slightly (between 3.0 and 3.7 for Rh and close to 2 for Ir). Kinetic and mechanistic studies have been carried out using these cationic complexes as catalyst precursors. For both complexes, the reaction proceeds according to the rate law ri = K1K2K3k4[M][olef][H2][CO]/([CO]2 + K1[H2][CO] + K1K2K3[olef][H2]). Both complexes react rapidly with CO to produce the corresponding tricarbonyl species [M(CO)3(PPh3)2]PF6, M = Rh, 2a; Ir, 2b, and with syn-gas to yield [MH2(CO)2(PPh3)2]PF6, M = Rh, 3a; Ir, 3b, which originate by CO dissociation the species [MH2(CO)(PPh3)2]PF6 entering the corresponding catalytic cycle. All the experimental data are consistent with a general mechanism in which the transfer of the hydride to a coordinated olefin promoted by an entering CO molecule is the rate-determining step of the catalytic cycle.  相似文献   

14.
The liquid-phase alkylation of phenol with 1-dodecene was carried out over WOx/ZrO2 solid acid catalysts. The catalysts were prepared by wet impregnation method using zirconium oxyhydroxide and ammonium metatungstate. Catalysts with different WO3 loading (5–30 wt.%) were prepared and calcined at 800 °C and catalyst with 15% WO3 was calcined from 700–850 °C. All the catalysts were characterized by surface area, XRD, and FTIR. The catalyst with 15% WO3 calcined at 800 °C (15 WZ-800) was found to be the most active in the reaction. The effect of temperature, molar ratio and catalyst weight on dodecene conversion and products selectivity was studied in detail. Under the optimized reaction conditions of 120 °C, phenol/1-dodecene molar ratio 2 and time 2 h, the catalyst 15 WZ-800 gave >99% dodecene conversion with 90% dodecylphenol selectivity. Comparison of the catalytic activity of 15 WZ-800 with sulfated zirconia calcined at 500 °C (SZ-500) and Hβ zeolite showed that activity of SZ-500 was lower than that of 15 WZ-800, while Hβ zeolite showed negligible activity. It is observed that the presence of water in the reaction mixture was detrimental to the catalytic activity of WOx/ZrO2. The catalyst 15 WZ-800 also found to be an efficient catalyst for alkylation of phenol with long-chain olefins like 1-octene and 1-decene.  相似文献   

15.
The 61Πu state of sodium dimer has been observed up to v = 53 in excitation spectra of the system, recorded by polarisation labelling spectroscopy technique. The Dunham coefficients are derived and the potential energy curve constructed by the inverted perturbation approach method. Equilibrium constants for the 61Πu state of Na2 are: Te = 35446.06 ± 0.04 cm−1 (with respect to the minimum of the electronic ground state), Y10 = 111.388 ± 0.019 cm−1, Y01 = 0.112122 ± 0.000017 cm−1.  相似文献   

16.
The samples of La0.4Sr0.6Co1−yFeyO3−δ (y = 0.2 and 0.4) were prepared using both conventional ceramic technique and nitrate–citrate precursors technique. The phase identification was made by X-ray diffraction method. The refinement of structural parameters from the XRD and neutron diffraction measurements was performed by full profile Rietveld analysis. Neutron diffraction showed that both samples possess distorted perovskite-type structure. Oxygen nonstoichiometry was measured by chemical analysis and thermogravimetry (TG) analysis in the range 20 ≤ T/°C ≤ 900 and 2E-5 ≤ pO2/atm ≤ 4E-1. TG-experiments indicate a relatively fast and reversible oxygen exchange at pO2 > 1E-2 atm. Mass saturation occurs at T < 300 °C upon cooling. The absolute value of oxygen nonstoichiometry was determined by iodometric titration measurements. It was found that both samples have practically stoichiometric composition at 300 °C in air and δ increases with increasing temperature and decreasing oxygen partial pressure.  相似文献   

17.
The mechanism of the SO2 + HO2 reaction was studied theoretically for the first time. Three product channels were revealed, namely, O2 + HOSO, O2 + HSO2, and OH + SO3. The O2 + HOSO channel dominates the reaction under combustion conditions. A five-member-ring complex [SO2–HO2] exists at the entrance of the reaction. The structure and binding energy (De and D0) of the SO2–HO2 complex have been calculated. In view of D0 = 21.2 ± 2.0 kJ mol−1, the SO2–HO2 complex should be stable at low temperature. The infrared spectra and frequency shifts were calculated for both SO2–HO2 and SO2–DO2, and compared with the available experimental data.  相似文献   

18.
Photoacoustic calorimetry (PAC) allows measurement of the energetics of reactive intermediates. Here, we report the examination of the metal carbonyl η5-CpMn(CO)3 (Cp, cyclopentadiene) via time-independent PAC, in a homologous series of solvents. The measured heat releases allow one to determine simultaneously the enthalpy and volume change resulting from the photodissociation of CpMn(CO)3. While the photoacoustic signal results from both of these processes, it has often been assumed that the volume change contribution to the observed photoacoustic signal is negligible for small molecules undergoing photodissociation. The current study tests the assumption of a negligible reaction volume by using a more complete treatment. The reaction of an equimolar number of photons and CpMn(CO)3 molecules, the subsequent photodissociation of the Mn–CO bond, and the ligation of a solvent molecule in an alkane solvent yields ΔHobs = 32.7 ± 0.7 kcal/mol and ΔVchem = 11.0 ± 1.3 mL/mol, both of which are independent of the quantum yield of photodissociation. A detailed analysis of the quantum yield is included (using both previously reported measurements, and new data from this work), from which we determine Φdiss = 0.635. This quantum yield allows us to determine ΔHrxn = 51.6 kcal/mol and ΔVrxn = 17.3 mL/mol. These results demonstrate that if the contribution of the reaction volume change to the photoacoustic signal is ignored, the reaction enthalpy derived would underestimate the true value by 7%. We also estimate the BDE{Cp(CO)2Mn–CO} to be 59.4 kcal/mol.  相似文献   

19.
Micellar-catalyzed alkaline hydrolysis of 2,4-dinitrochlorobenzene (DNCB) in the presence of a conventional cationic surfactant CTAB or a cationic gemini surfactant 1,2-ethane bis(dimethyldodecylammonium bromide) (12-2-12) were studied spectrophotometrically at 25 °C. It was found that both CTAB and 12-2-12 micelles accelerated the alkaline hydrolysis of DNCB, and the binding constant of the substrate to the micelle, KS, for 12-2-12 (KS = 310 M−1) was larger than that for CTAB (85 M−1), which suggested that DNCB molecules bound with gemini micelles more easily than with CTAB. However, the second-order rate constant in micellar pseudophase (kM = 1.22 × 10−3 s−1) for 12-2-12 was lower than kM for CTAB (4.01 × 10−3 s−1) because the substrate may enter the interior of the 12-2-12 micelles. It was found also that 12-2-12 had a similar catalysis mechanism to CTAB when the concentration of 12-2-12 was relatively low (ca. <5 mM). However, above this concentration, higher microviscosity and significant increases in aggregation number and micelle size with increased surfactant concentration may remarkably influence the hydrolysis reaction.  相似文献   

20.
In a constant ionic medium, corresponding to a physiological environment (Ic = 0.15 mol dm−3 NaCl), the solubilities of anhydrous uric acid, uric acid dihydrate and monosodium urate monohydrate have been measured as a function of p[H] = −log[H+](2-8) and temperature (25°, 32°, 37° and 42°C). The solubility equilibria in the uric acid-sodium urate-water system are discussed on the basis of the solubility constants (Ks) and the first dissociation constant (K1) of uric acid and the solubility product (Ks0) of monosodium urate. The quantities measured in this work are in good agreement with literature values, however, the present solubility data have a much higher precision.  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