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 15 毫秒
1.
Carbon-13 fractionation observed in the course of carbon monoxide formation in the reaction of phenylacetylene with the large excess of liquid formic acid in the temperature interval 20–100°C has been investigated and compared with the13C fractionation in the dehydration of pure liquid formic acid. The anomalous temperature dependence of the13C fractionation has been interpreted as caused by the change of the kinetics and of the mechanism of CO formation from the one involving13C–H bond rupture rate determining step (operating in the presence of phenylacetylene) to the mechanism according to which HCOOH decarbonylates in liquid state. No large increase of the13C fractionation with rising of the reaction temperature from 70 to 134°C has been found in the case of decarbonylation of F.A. in the presence of large excess of phenylacetylene. The13C KIE was of 1.020 in the temperature interval 90–133.7°C in this case.  相似文献   

2.
The13C fractionation has been studied in the reaction of phenylacetylene with the excess of liquid Merck formic acid at 30 and 40 °C to see the contribution of the13C fractionation in the formolysis of transient -formoxystyrene to the experimentally observed global13C fractionation. The13C fractionation has been investigated also in the hydration of 1 ml of PhCCH with 1 ml of formic acid in the temperature interval of 80–100°C. The13C KIE equal to 1.0168 at 91.75 °C and 1.0167 at 100°C indicate that the self-decomposition of formic acid in such experimental conditions is already largely suppressed. The isotope effect is discussed within the framework of the sequence of reaction steps leading to acetophenone and carbon monoxide production listed in part I.  相似文献   

3.
The13C kinetic isotope effect (K.I.E.) in the decarbonylation of formic acid of natural isotopic composition in 85% orthophosphoric acid, in 100% H3PO4, and in pyrophosphoric acid has been measured in different temperature intervals ranging from 19 to 133 °C. In 85% H3PO4 the carbon-13 K.I.E. is determined by the fractionation of carbon isotopes expected for C–O bond rupture (k 12/k 13=1.0531 at 70°C). In 100% H3PO4 the13C K.I.E. indicates that C–H bond rupture is the major component of the reaction coordinate motion (thek 12/k 13 lay in the range of 1.026–1.017 over the range 30–70 °C). In pyrophosphoric acid the fractionation factor for13C equals 1.010 at 19 °C. Activation parameters for the decarbonylation of H12COOH in phosphoric acid media have been determined also and suggestions concerning the intimate mechanisms of decarbonylation of formic acid in dilute and concentrated phosphoric acids are made.  相似文献   

4.
Carbon-13 intramolecular kinetic isotope effects in the decarbonylation of oxalic acid dihydrate of natural isotopic composition by SO3 and by fuming sulphuric acid at room temperature and decarbonylation of oxalic acid dihydrate by 100% H3PO4 in the temperature interval 80–150°C have been determined. The obtained isotopic and kinetic results have been compared with the earlier13C experimental and theoretical studies in other solvents.  相似文献   

5.
The13C kinetic isotope effect fractionation in the decarbonylation of lactic acid (LA) of natural isotopic composition by concentrated phosphpric acids (PA) and by 85% H3PO4 has been studied in the temperature interval of 60–150°C. The values of the13C(1) isotope effects in the decarbonylation of lactic acid in 100% H3PO4, in pyrophosphoric acid and in more concentrated phosphoric acids are intermediate between the values calculated assuming that the C(1)–OH bond is broken in the rate-controllin gstep of dehydration and those calculated for rupture of the carbon-carbon bond in the transition state. In the temperature interval of 90–130°C the experimental13C fractionation factors determined in concentrated PA approach quite closely the13C fractionation corresponding to C(2)–C(1) bond scission. the13C(1) kinetic isotope effects in the decarbonylation of LA in 85% orthophosphoric acid in the temperature range of 110–150°C coincide with the13C isotope effects calculated assuming that the frequency corresponding to the C(1)–OH vibration is lost in the transition state of decarbonylation. A change of the mechanism of decarbonylation of LA in going from concentrated PA medium to 85% H3PO4 has been suggested. A possible secondary18O and a primary18O kinetic isotope effect in decarbonylation of lactic acid in phosphoric acids media have been discussed, too.  相似文献   

6.
Pyrolysis of t-butyl formate, (CH3)3C-O-CHO, has been carried out in a carrier gas stream of Ar or N2 in a temperature range of 200–400°C. Between 200 and 300°C, the pyrolysis yielded an equimolar mixture of HCOOH and (CH3)2C=CH2. The results have been used as a calibration method for determining the concentration of the gas-phase HCOOH monomer without interference from the formation of the formic acid dimer. Using this technique, the gas-phase infrared absorption cross-section of HCOOH at 1105 cm–1 (peak to valley) for the resolution of 0.5 cm–1 has been determined to be 6.76×10–19 cm2 molecule–1.  相似文献   

7.
The isotopic composition of the consecutive fractions of carbon monoxide produced in the decarbonylation of liquid formic acid of natural isotopic composition initiated by addition of phosphorus pentoxide has been measured in the temperature interval 19–100°C and the observed gradual decrease of the PDB values and the increase of thek 12/k 13 ratio of the isotopic specific rate constants (KIE values) for each next fraction of CO have been interpreted in terms of conclusions presented in the first paper from this series1 concerning the decarbonylation of HCOOH (F.A.) in concentrated and diluted with water phosphoric acid media. The initial fast dehydration of F.A. by phosphoric anhydride, P2O5, proceeds at room temperture with about 1% carbon-13 KIE. The (k 12/k 13) values increase with time, as the decarbonylation slows down due to the hydration of phosphorus pentoxide with water generated in dehydration of HCOOH and reach the plateau values characteristic for each reaction temperature. These increasing very slowly with reaction times at intermediate temperatures maximum values of (k 12/k 13) ratios are quite close to values of13C KIE observed in the decarbonylation of pure F.A. (k 12/k 13=1.0443 at 81°C). Addition of water to liquid F.A. at 90°C and at 100°C caused the further increase of the13C KIE. The detailed discussion of the13C KIE in the HCOOH–P2O5 system has been given.  相似文献   

8.
The values of the second dissociation constant, pK2, and related thermodynamic quantities of 4-(N-morpholino)butanesulfonic acid (MOBS) and N-tris(hydroxymethyl)-4-aminobutanesulfonic acid (TABS) have already been reported over the temperature range 5–55°C including 37{°}C. This paper reports the pH values of twelve equimolal buffer solutions at designated pH (s) with the following compositions: (a) mixtures of MOBS (0.05 mol-kg–1) + NaMOBS (0.05 mol-kg–1); (b) MOBS (0.08 mol-kg–1) + NaMOBS (0.08 mol-kg–1); (c) MOBS (0.08 mol-kg–1) + NaMOBS (0.08 mol-kg–1) + NaCl (0.08 mol-kg–1); (d) TABS (0.05 mol-kg–1) + NaTABS (0.05 mol-kg–1); and (e) TABS (0.08 mol-kg–1) + NaTABS (0.08 mol-kg–1); and (f) TABS (0.08 mol-kg–1) + NaTABS (0.08 mol-kg–1) + NaCl (0.08 mol-kg–1). Two buffer solutions have ionic strengths I= 0.05 mol-kg–1, another two have I=0.08 mol-kg–1, and the remaining two buffer solutions have I= 0.16 mol-kg–1, which is close to that of the clinical fluids (blood serum). These buffers have been recommended as a useful pH standard for the measurements of physiological solutions. Conventional pH values of all six buffer solutions from 5–55°C, as well as those obtained from the liquid junction potential correction at 25 and 37{°}C have been calculated. The flowing-junction calomel cell has been utilized to measure Ej, the liquid junction potential.  相似文献   

9.
The dissociation quotients of formic acid were measured potentiometrically from 25 to 200°C in NaCl solutions at ionic strengths of 0.1, 0.3 1.0, 3.0, and 5.0 mol-kg–1. The experiments were carried out in a concentration cell with hydrogen electrodes. The resulting molal acid dissociation quotients for formic acid, as well as a set of infinite dilution literature values and a calorimetrically-determined enthalpy of reaction, were fitted by an empirical equation involving an extended Debye Hückel term and seven adjustable parameters involving functions of temperature and ionic strength. This regressional analysis yielded the following thermodynamic quantities for 25°C: logK=–3.755±0.002, Ho=–0.09±0.15 kJ-mol–1, So=–72.2±0.5 J-K–1-mol–1, and C p o =–147±4 J-K–1-mol–1. The isocoulombic form of the equilibrium constant is recommended for extrapolation to higher temperatures.  相似文献   

10.
Membranes, based on tri-n-octylamine (TOA) xylene liquid, supported in hydrophobic microporous films have been used to study the transport of Pd(II) ions, after extraction into the membrane. Various parameters, such as the effect of hydrochloric acid concentration in the feed solution, TOA concentration in the membrane phase, effect of stripping agent like nitric acid concentration, and temperature on the flux of Pd(II) ions across the liquid membranes have been investigated. The optimum conditions of transport for these metal ions determined are, TOA concentration, 1.25 mol·dm–3, HCl concentration in the feed solution, 5 mol·dm–3, and concentration of nitric acid used as a stripping, agent 5 mol·dm–3. The maximum values of the flux and permeability determined under the optimum condition are 23·10–6 mol·m–2·s–1 and 2.40·103 m2·s–1 at 25°C. The results obtained have been used to elucidate the mechanism of palladium transport.  相似文献   

11.
High precision densities of sodium chloride solutions at a constant pressure of 200 bar and temperatures between 175°C and 350°C have been measured by a mercury displacement technique. The densities have been converted to apparent molar volumes. The apparent molar volumes decrease with increasing temperature and decreasing concentration whereas the concentration effect increases with temperature. Standard partial molar volumes range from 8.0 cm3-mol–1 at 175°C to –600 cm3-mol–1 at 350°C. The results indicate the applicability of the unextended Debye-Hückel limiting law up to concentrations of 0.02 mol-kg–1.  相似文献   

12.
Raman spectra of gold bromide complexes in acidic solutions (pH=–0.3–3) have been recorded at 25° to 300°C and at pressures on the liquid vapor curve for the system. At 25°C, only the square planar Au(III) bromide complex, AuBr 4 , is present in solution with bands at approximately 105, 197 and 215 cm–1. However, in these acidic solutions, when the temperature is 50°C or higher, the square planar Au(III) bromide complex is partially transformed into the linear Au(I) bromide complex, AuBr 2 , with a single band near 208 cm–1. The transformation of the Au(III) square planar tetrabromo complex into the Au(I) linear dibromo complex is also favored by a reduction of the oxygen fugacity and an increase in pH.  相似文献   

13.
The second dissociation constant pK2 of 3-(N-morpholino)propanesulfonic acid (MOPS) has been determined at eight temperatures from 5 to 55°C by measurements of the emf of cells without liquid junction, utilizing hydrogen electrodes and silver–silver chloride electrodes. The pK2 has a value of 7.18 ± 0.001 at 25°C and 7.044 ± 0.002 at 37°C. The thermodynamic quantities G°, H°, S°, and C p o have been derived from the temperature coefficients of the pK 2. This buffer at ionic strength I = 0.16 mol-kg–1 close to that of blood serum, has been recommended as a useful secondary pH standard for measurements of physiological fluids. Five buffer solutions with the following compositions were prepared: (a) equimolal mixture of MOPS (0.05 mol-kg–1) + NaMOPS, (0.05 mol-kg–1); (b( MOPS (0.05 mol-kg–1) + NaMOPS (0.05 mol-kg–1) + NaCl (0.05 mol-kg–1); (c) MOPS (0.05 mol-kg–1) + NaMOPS (0.05 mol-kg–1); + NaCl (0.11mol-kg–1); (d) MOPS (0.08 mol-kg–1) + NaMOPS (0.08 mol-kg–1); and (e)MOPS (0.08 mol-kg–1) + NaMOPS (0.08 mol-kg–1) + NaCl (0.08 mol-kg–1).The pH values obtained by using the pH meter + glass electrode assembly are compared with those measured from a flow–junction calomel cell saturated with KCl (cell B), as well as those obtained from cell (A) without liquid junction at 25 and 37°C. The conventional values of the liquid junction potentials E j have been obtained at 25 and 37°C for the physiological phosphate reference solution as well as for the MOPS buffers (d) and (e) mentioned above.  相似文献   

14.
The possibilities of liquid membrane preconcentration of neptunium from environmental samples of different nature have been studied. The use the solid-supported liquid membrane containing a trioctylmethylammonium nitrate carrier allows to achieve preconcentration factors of 102–5×102. The teflon solid support does not interact with the luminescent matrix (CaF2, PbMoO4) during calcination at 900 °C, so it makes practical to measure the neptunium content by luminescence without reextraction to aqueous solution. As a result, the detection limit of neptunium is lowed down to 10–13 g ml–1 and 5×10–13 g g–1 for pure solutions and soils, respectively.  相似文献   

15.
The kinetic behavior of deuteriation of benzoic acid with D2O acidified with HCl in the presence of homogeneous K2PtCl4 catalyst has been investigated in the 100–130°C temperature interval. The quasiunimolecular H/D rate constants at 100 and 130°C corresponding to an exchange process in ortho positions of the substituted benzene ring hydrogens were determined by1H NMR integration signal. These same constants for meta and para positions have been deduced by analysis of the composite1H NMR signal, and the Arrhenius activation energies for these exchange reactions were estimated.  相似文献   

16.
Depth profiles of Ga2O3/a-SiO2/Al2O3- substrate, Ga2O3/a-Si3N4/Al2O3- substrate, and Ga2O3/Al2O3 substrate thin layers were determined by the SNMS/HFM method. Al diffusion from the Al2O3 substrate was investigated after 50, and in some cases after 600 hours of heat treatment time at different temperatures (600 °C,850 °C,950 °C,1050 °C and 1150 °C). The diffusion coefficient of Al at 850 °C was found to be D Al=8.7 * 10–18 cm2/s in amorphous SiO2; D Al=1.5*10–17 cm2/s in amorphous Si3N4 and D Al=5.5* 10–16 cm2/s in Ga2O3 at 600 °C, respectively. The possible diffusion mechanism is explained in terms of the metal-oxygen bond-strengths. Although the studied materials have high resistivity at room temperature, the applied SNMS/HFM method has proven to be an efficient surface analytical tool even in these cases.Dedicated to Professor Dr. rer. nat. Dr. h.c. Hubertus Nickel on the occasion of his 65th birthday  相似文献   

17.
Simple two-parameter Hückel and Pitzer equations were used for the calculation of the activity coefficients of aqueous hydrochloric acid at temperatures 0–60°C up to a molality of 2.0 mol-kg–1. The data obtained by Harned and Ehlers(2,3) on galvanic cells without a liquid junction were used in the parameter estimations of these equations. These data consist of sets of measurements at the temperature intervals of 5°C. It was observed that all estimated parameters follow very simple equations with respect to temperature. They are either constant or depend linearly on the temperature. The values for the activity coefficient parameters calculated by these simple equations are recommended here. The recommended parameter values were tested by predicting the data of Gupta, Hills, and Ives,(5) consisting of cell measurements from 5 to 45°C and molalities up to 1.0 mol-kg–1, and the data of Bates and Bower,(4) which extend to 95°C but measurements were only made on molalities less than about 0.1 mol-kg–1. The activity coefficients obtained by the new equations were also compared to those calculated by the Pitzer equations with the parameter values determined by Saluja, Pitzer, and Phutela(6) from calorimetric data. The agreement observed was excellent up to a molality of 1.5 mol-kg–1 at temperatures from 0 to 60°C.  相似文献   

18.
We have investigated the proton conductivities of the sol-gel-derived P2O5-SiO2 glass at –50 to 120°C. The obtained glass is porous, where the surface area, pore volume and pore diameter are 740 m2/g, 0.5 cm3/g and <5 nm, respectively. The freezing temperature of water molecules adsorbed in the pores was –20°C, which is much lower than that of free liquid water due to the quantum size effect of the water confined in the pores. The electrical conductivities followed the Arrhenius equation in the temperatures between –20 and 120°C. Below –20°C, the adsorbed-water molecules were frozen, resulting in a rapid decrease of the proton conductivity. Considering the high conductivity, chemical and thermal stability, this oxide glass membranes have potential for the fuel cell membrane.  相似文献   

19.
Study of the extraction of W(VI) ions using supported liquid membrane has been carried out. The carrier used for this metal ion transport, is tri-n-octylamine (TOA) dissolved in xylene. The liquid was supported in microporous polypropylene film. The parameters studied are effect of carrier concentration in the membrane, acid concentrations in the feed solution, concentration of stripping agent on transport of W(VI) ions and of temperature on the transport properties of these supported liquid membranes. The optimum conditions of transport for these metal ions determined are, TOA concentration, 0.66 mol·dm–3 (TOA); HF concentration in the feed solution, 0.01 mol·dm–3 and concentration of NaOH used as stripping agent 2.5 mol·dm–3. The maximum flux and permeability determined under optimum conditions are 3.06·10–5 mol·m–2·s–1 and 8.44·10–11 mol· ·m2·s–1 at 25±2°C and 4.21·10–5 mol·m–2·s–1 and 11.55·10–11 mol·m2·s–1 at 65°C, respectively. The diffusion coefficients for the metal ion carrier complex in the membrane have also been determined. Under the optimum conditions the value for the metal ion carrier complex is 0.14·10–11 mol·m2·s–1. Mechanism of transport and the complex formed in the presence of HF have also been discussed. The transport process involves two carrier amine molecules and two protons.  相似文献   

20.
The kinetics of oxidation of diethylene glycol, triethylene glycol, and polyethylene glycols (PEGs) with molecular weights ranging from 400 to 2000 in the presence of Cu(II) ions and bases was studied. It was found that ethylene glycols can be oxidized by molecular oxygen in anhydrous media in a temperature range of 30–60°C at anomalouosly high rates which are higher than the rates of chain-radical PEG autoxidation by several orders of magnitude. Only terminal hydroxyl groups were subjected to oxidation. The reaction occurs with the cleavage of a C–C bond and results in the formation of formic acid and a PEG with the number of –(CH2CH2O)– groups lower than that in the parent compound by unity. The rate and selectivity of PEG oxidation were found to strongly depend on the molecular weight of the polymer; from diethylene glycol to PEG 2000, the specific rate of oxidation increased by a factor of 60 in terms of terminal hydroxyl groups. An oxidation mechanism was suggested, which involves the formation of ternary complexes [Cu2+···A···O2], which undergo further degradation by a many-electron concerted mechanism to form formic acid and, probably, an unstable hemiacetal {RO–CH2OH}. The rapid oxidative degradation of the latter leads to the formation of PEG with a lower molecular weight.  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