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.
Deforming tubing (e.g., coiling) reduces axial dispersion in flow-injection systems and post-column reactors. An improvement is presented for a previous theoretical treatment of axial dispersion in coiled tubes. New experimental data are also presented for different coil radii. Changes were necessary because smaller-than-predicted dispersion improvements were observed. This contradicts earlier treatments that predict a linear dependence on the aspect ratio of the coil. Evidence is presented that this is explained by assuming a boundary layer thickness much smaller than the radius of the tube and proportional to the ratio of radial velocities. This is a special case of earlier theory and it predicts an almost curvature-independent change in dispersion. The inclusion of the critical Reynolds number, Rec, to form a reduced parameter, (Re/Rec), accounts for all observations, including some previously-defined empirical constants. The dispersion is found to increase as (Re/Rec)2/3Sc at low fluid velocities and to decrease as (Re1/6c/(Re/Rec)4/3Sc0.08) at higher velocities. Experiments are reported for several coil radii and with a range of proteins and protein mixtures.  相似文献   

2.
t-Butyl methyl ether has been UV photolysed (λ = 185 nm) to a maximal conversion of less than 0·1%. A study of the products (quantum yields) has been made: methanol (0·405), t-butanol (0·20), isobutene (0·178), t-butyl neopentyl ether (0·142), t-butyl ethyl ether (0·134), 1,2-di-t-butoxyethane (0·097), methane (0·056), isobutane (0·046), isopropenyl methyl ether (0·030), hydrogen (0·020), neopentane (0·016), ethane (0·015), formaldehyde (0·012), 2-methoxy-2-methyl-4-t-butoxybutane (0·005), hexamethylethane (0·0048), 2-methoxy-2-methylbutane (0·0027), 2-methoxy-2-methyl-3-t-butoxypropane (0·002), isopropyl methyl ether (0·0015), formaldehyde t-butyl methyl acetal (0·001), formaldehyde di-t-butyl acetal (0·001), 2-methoxy-2-methyl-4,4-dimethylpentane (0-001), 2-methoxy-2-methyl-3,3-dimethylbutane (0·0003), 2,5-dimethoxy-2,5-dimethylhexane (0·0002), di-t-butyl ether (5 · 10?5), 2,2-dimethyloxirane (?, <- 0·001). There is no decomposition of the t-BuO radical into acetone (< 5 · 10?4) and CH3. Cyclisation reactions leading to α,α-dimethyloxetane (< 10?4) and 1-methoxy-1-methylcyclopropane (< 10?4) do not occur. The material balance yields C5H11·97O1·018.The main modes of fragmentation (ca 82%) are represented by the homolytic CO bond split, either into t-butyl and methoxy (ca 52%) or into t-butoxy and methyl (ca 30%), Fragmentation into methanol and isobutene (8·5%) as well as into formaldehyde and isobutane (2%) are further modes of decomposition. The break of a CC linkage (4·5%) mainly occurs by elimination of molecular methane. A CH bond split has a probability of ca 3% with the methoxy CH bond the more likely one to break.  相似文献   

3.
The kinetics of the gas phase pyrolysis were investigated for 2-acetoxy-4-chloropentane, the simplest model for the macromolecule of VC-VAc copolymer; both manometric and GLC method were used. It was concluded that the elimination of the first acid molecule from the model is a homogeneous unimolecular first order reaction with rate constants within the range 310–350° given by: log k1 = (13·14 ± 0·17)—(44·9 ± 0·5)/2·303 RT for the iso and log k1 = (13·23 + 0·15)—(44·9 ± 0·4)/2·303 RT for the syndio isomer. (Frequency factors in sec?1 units, activation energies in kcal mol?1.)Application of the results to the copolymer leads to the conclusion that the pronounced minimum observed by some authors in the plot of thermal stability against copolymer composition cannot be adequately explained solely in terms of the low thermal stability of an alternating unit.  相似文献   

4.
The gas phase pyrolyses of cis-5-chloro-3-heptene (in the range 267–337°) and cis- and trans-5-acetoxy-3-heptene (300–378°) are homogeneous unimolecular first-order reactions with rate constants given respectively by: log k = (12·03 = 0·13) - (36·2 ± 0·4)/2·303 RT and (12·80 ± 0·11) - (43·0 ± 0·3)/2·303 RT. (Frequency factors in sec?1 units, activation energies in kcal mol?1.) No significant differences were found between the rates of decomposition of cis and trans isomers of 5-acetoxy-3-heptene. From the decomposition of these models of poly(vinyl chloride) and poly(vinyl acetate), some conclusions about the role of internal unsaturated groups in the thermal decomposition of both polymers were drawn. The possibility that groups with cis internal double bonds are the most labile structures in a poly(vinyl) chloride macromolecule is discussed.  相似文献   

5.
Ionic conductivity of double vanadate Ag3Sc2(VO4)3 with the NASICON structure is studied by the method of impedance spectroscopy in the frequency range from 5 to 5 × 105 Hz and in the temperature range of 300–827 K. The vanadate Ag3Sc2(VO4)3 is prepared in the form of fine crystalline powder by solid-state synthesis from V2O5, Sc2O3, and AgNO3 at 1173 K. The conductivity of Ag3Sc2(VO4)3 ceramic samples σ = 8 × 10?3 S/cm (at 563 K). The σ vs. T curve demonstrates an anomaly at 563–623 K which corresponds to thermal effects in this temperature range. The values of enthalpy of ion transport activation are ΔH = 0.40 ± 0.05 eV (T < 563 K) and ΔH = 0.30 ± 0.05 eV (T > 623 K). The ionic conductivity of Ag3Sc2(VO4)3 is due to Ag+ ions localized in channels of the framework structure of the NASICON type.  相似文献   

6.
From extraction experiments and γ-activity measurements, the exchange extraction constant corresponding to the equilibrium Ba2+(aq) + Sr2+(nb) ?1·Ba2+(nb) + Sr2+(aq) taking place in the two-phase water–nitrobenzene system (1 = beauvericin; aq = aqueous phase, nb = nitrobenzene phase) was evaluated as log K ex (Ba2+, Sr2+) = 1.2 ± 0.1. Further, the stability constant of the beauvericin–barium complex (abbrev. Ba2+) in nitrobenzene saturated with water was calculated for a temperature of 25 °C: log β nb (Ba2+) = 9.5 ± 0.2. Finally, by using quantum mechanical DFT calculations, the most probable structure of the Ba2+ complex species was predicted.  相似文献   

7.
The reaction of Ca(BH4)2 · 1.5Dg with decaborane-14 in diglyme at 85°C yields Ca(B11H14)2 · 4Dg. The duration of the reaction is 20 h. The molar ratio of Ca(BH4)2 · 1.5Dg: B10H14 is 1: 3.5. The yield is 84.7%. The synthesized calcium tetradecahydroundecaborate Ca(B11H14)2 · 4Dg is a stable compound soluble in water and a number of organic solvents. The study of the compound was carried out using elemental analysis and IR and NMR 11B {1H} spectroscopy. Calcium undecaborate can be used for synthesizing other salts with boron hydride nido-borates.  相似文献   

8.
The enthalpy change of the reaction at 298 K between Br2 (l) and Sn(c) in CS2 as solvent giving SnBr4 (s) has been determined by calorimetry to be (?374, 2±1.4) kJ·mol?1, [(?89.45±0.33) kcal·mol?1]. By the same method the heat of solution of SnBr4 (c) in CS2 has been found to be (11.9±0.3) kJ·mol?1, [(2.84±0.08) kcal·mol?1]. Combining these results, a value of (?386.1±1.5) kJ·mol?1, [(?92.3±0.4) kcal·mol?1] is derived for the standard heat of formation of SnBr4 (c). Substituting this figure in the thermochemical cycle hitherto used for calculating the heat of formation of SnBr4 (c) gives ?124.3 kcal·mol?1 for the standard heat of formation of SnCl4 (l), which is in reasonable agreement with a recent determination of this quantity8.  相似文献   

9.
The thermal unimolecular decomposition of three vinylethers has been studied in a VLPP apparatus. The high-pressure rate constant for the retro-ene reaction of ethylvinylether was fit by log k (sec?1) = (11.47 + 0.25) - (43.4 ± 1.0)/2.303 RT at <T> = 900 K and that of t - butylvinylether by log k (sec?1) = (12.00 ± 0.27) - (38.4 ± 1.0)/2.303 RT at <T> = 800 K. No evidence for the competition of the higher energy homolytic bond-fission process could be obtained from the experimental data. The rate constant compatible with the C? O bond scission reaction in the case of benzylvinylether was log k (sec?1) = (16.63 ± 0.30) - (53.74 ± 1.0)/2.303 RT at <T> = 750 K. Together with ΔHf,3000(benzyl·) = 47.0 kcal/mol, the activation energy for this reaction results in ΔHf,3000(CH2CHO) = +3.0 ± 2.0 kcal/mol and a corresponding resonance stabilization energy of 3.2 ± 2.0 kcal/mol for 2-ethanalyl radical.  相似文献   

10.
Excess molar volumes VmE at 298.15 K were obtained, as a function of mole fraction x, for series I: {x1-C4H9Cl + (1 ? x)n-ClH2l + 2}, and II: {x1,4-C4H8Cl2 + (1 ? x)n-ClH2l + 2}, for l = 7, 10, and 14. 10, and 14. The instrument used was a vibrating-tube densimeter. For the same mixtures at the same temperature, a Picker flow calorimeter was used to measure excess molar heat capacities Cp, mE at constant pressure. VmE is positive for all mixtures in series I: at x = 0.5, VmE/(cm3 · mol?1) is 0.277 for l = 7, 0.388 for l = 10, and 0.411 for l = 14. For series II, VmE of {x1,4-C4H8Cl2 + (1 ? x)n-C7H16} is small and S-shaped, the maximum being situated at xmax = 0.178 with VmE(xmax)/(cm3 · mvl?1) = 0.095, and the minimum is at xmin = 0.772 with VmE(xmin)/(cm3 · mol?1) = ?0.087. The excess volumes of the other mixtures are all positive and fairly large: at x = 0.5, VmE/(cm3 · mol?1) is 0.458 for l = 10, and 0.771 for l = 14. The Cp, mEs of series I are all negative and |Cp, mE| increases with increasing l: at x = 0.5, Cp, mE/(J · K?1 · mol?1) is ?0.56 for l = 7, ?1.39 for l = 10, and ?3.12 for l = 14. Two minima are observed for Cp, mE of {x1,4-C4H8Cl2 + (1 ? x)n-C7H16}. The more prominent minimum is situated at xmin = 0.184 with Cp, mE(xmin)/(J · K?1 · mol?1) = ?0.62, and the less prominent at xmin = 0.703 with Cp, mE(xmin)/(J · K?1 · mol?1) = ?0.29. Each of the remaining two mixtures (l = 10 and 14) has a pronounced minimum at low mole fraction (xmin = 0.222 and 0.312, respectively) and a broad shoulder around x = 0.7.  相似文献   

11.
From extraction experiments and γ-activity measurements, the exchange extraction constant corresponding to the equilibrium Ca2+ (aq)+1·Sr2+ (nb) ⇆ 1·Ca2+ (nb) + Sr2+ (aq) taking place in the two-phase water-nitrobenzene system (1 = valinomycin; aq = aqueous phase, nb = nitrobenzene phase) was evaluated as log K ex (Ca2+, 1·Sr2+) = 2.4±0.1. Further, the stability constant of the valinomycin-calcium complex (abbrev. 1·Ca2+) in nitrobenzene saturated with water was calculated for a temperature of 25 °C: log β nb (1·Ca2+) = 8.3±0.1. By using quantum mechanical DFT calculations, the most probable structure of the 1·Ca2+·2H2O complex species was predicted. In this complex, the “central” Ca2+ cation is bound by strong bonds to two oxygen atoms of the respective water molecules and to four ester carbonyl oxygens of the parent valinomycin ligand 1. Finally, the calculated binding energy of the considered complex 1·Ca2+·2H2O is −319.2 kcal/mol, which confirms the relatively high stability of this cationic complex species.  相似文献   

12.
Y. Tanaka  S.R. Velen  S.I. Miller 《Tetrahedron》1973,29(21):3271-3283
About twenty new H-1,2,3-triazoles (T) were readily synthesized by nucleophilic attack of sodium azide on activated acetylenes in dimethylformamide. Typical activating groups were COR, COOR, O2NC6H4 PO(OC2H5, COT, and (C6H5)3P+. Propynyl 4-triazolyl ketone or phenylethynyl 4-triazolyl ketone may be converted to acylic adducts (triazolylketoenamines), biheteroaromatic systems (isoxazolytriazoles, pyrazolytriazoles), as well as to ditriazolyl ketones. Certain T properties were examined in detail. The apparent pK′s for our group of ca 30 triazoles were in the range 4·95?9·45 in ethanol-water (v/v 1/1) at 25°. The Hammett correlation for five 4-aryl-T was log Ka = 0·89σ? ?9·21 and for seven 4-aryl-5-carbethoxy-T was log Ka = 1·45σ?6·95. The UV spectra of T are similar to those of other heteroaromatic and phenyl compounds: interesting analogies between triazolyl and phenyl, e.g., ”ortho“ crowding effects, appear to be indicated in the spectra of compounds related to biphenyl, stilhene and benzophenone. With regard to structure assignment on the basis of spectra, characteristic features of UV and IR spectra of the H-1,2,3-triazoles are discussed.  相似文献   

13.
The new scandium/aluminium co-doped NASICON phases Li1?+?x Al y Sc x???y Ti2???x (PO4)3 (x?=?0.3, y?=?0,0.1,0.2,0.3) were prepared by mechanical milling followed by annealing of the mixtures at 950 °C. X-ray diffraction of all samples showed the formation of NASICON structure with space group R-3c along with a minor impurity. Rietveld refinement of the X-ray data was performed to identify the structural variation. Doping with Sc3+ caused elongation of a- and c- axes for all the compounds when compared with undoped LiTi2(PO4)3. The compound Li1.3Sc0.3Ti1.7(PO4)3 showed a maximum of a?=?8.5504(7), c?=?20.986(3) Å at room temperature and exhibited highest coefficient of thermal expansion. The highest ionic conductivity (σ), 7.28×10?4 S cm?1 was observed for Li1.3Sc0.3Ti1.7(PO4)3, two orders of magnitude higher than for the undoped phase.  相似文献   

14.
Perchlorates of Sc3+, Y3+, La3+, and Eu3+ have each been extracted from 0.1 mol·dem?3 aqueous solution of μ□=1 with trioctylphosphine oxide (TOPO) in hexane. The hydration number of the extracted salts has been determined by Karl Fischer titration. Sc3+, Y3+ and EU3+ are extracted as the tetra- and hexa-solvates of TOPO but the assumption of octa-solvate on addition to the tetra-solvate explains the extraction data of La3+ well. The hydration number of tetra-solvates is 2 (Sc3+ and Eu3+), 3 (Y3+) and probably 4 for La3+ ion. That of hexa- and octa-solvates is 4–5.  相似文献   

15.
The reaction of Nd(BH4)3 · 3THF with decaborane-14 in diglyme at 85–90°C yields Nd(B11H14)3 · 4Dg. The duration of the reaction is 20 h. The molar ratio of Nd(BH4)3 · 3THF to B10H14 is is 1 :3.5. The product is precipitated with heptane from a diglyme solution. The yield is 70%. In an inert atmosphere, Nd(B11H14)3 · 4Dg is stable to 150°C and decomposes with an exotherm at 160–190°C. The IR spectrum of Nd(B11H14)3 · 4Dg in the region of B-H stretching vibrations contains an intense band at 2530 cm?1. The 11B {1H} NMR spectra of the synthesized compound in diglyme solutions contain signals of the tetradecahydro-nido-undecaborate anion B11H 14 ? (δ = -14.0, -15.6, and -16.5 ppm).  相似文献   

16.
The decompositions of the following poly(vinylacetate) models were investigated in the gas phase: 2-acetoxypentane (in the range 323–380°), meso and racem. 2,4-diacetoxypentane (325–375°), 4-acetoxy-1-pentene (302–375°), 4-acetoxy-2-pentene (314–381°), 5-acetoxy-2-hexene (328–391°), and 7-acetoxy-3,5-nonadiene (299–361°). The decompositions are homogeneous unimolecular first-order reactions with rate constants of 1·04 × 1013 exp(?45·/RT, 1·63 × 1013 exp(?45·4/RT), 1·01 × 1013 exp(?44·5/RT), 5sd76 × 1012 exp(?42·8/RT), 2·01 × 1013 exp(?45·5/RT), and 3·29 × 1012 exp(?41·5/RT) sec?1, respectively. The two 2,4-diacetoxypentane stereoisomers decompose at almost the same rate. From the course of decomposition of models in the gas phase, possible conclusions about decomposition of polymer are drawn. The role of configuration of sequences, double bonds, their number and geometric isomerism in the macromolecular chains are discussed in connection with elimination of acetic acid from poly(vinyl acetate).  相似文献   

17.
Extraction of vanadium(V) with 8-quinolinol into chlorobenzene is discussed. Three dimeric species are shown to be responsible for the extraction: 2VO3- + 4(HOx)o α (V2O3Ox4)o + 2OH-; log Kex,1 = -1.60 ± 0.10 2VO3- + 4(HOx)o + H+ + ClO4- α (V2O3H(Ox)4 · ClO4)o + 2OH-; log Kex,2 = 1.55 ± 0.10 2VO3- + 4(HOx)o + 2H+ + 2ClO4- α (V2O2Ox4 · 2ClO4)o + 2OH-; log Kex,3 = 2.65 ± 0.10 The vanadium(V) complex of 8-quinolinol has also been studied by thermogravimetry and i.r. and visible spectroscopy; an oxo-bridged dimeric structure is postulated. In contrast to 8-quinolinol, 2-methyl-8-quinolinol gives a monomeric vanadium(V) complex under the usual experimental conditions.  相似文献   

18.
A comparison was made between the experimental and the calculated dependence of the equilibrium potential on the ratio of the diffusion currents (E ? log(iK,k/iD,a)) for various combinations of intermediates. In addition the observed equilibrium potential was correlated to the calculated ratio log([FMN]/[FMNH2]). A good correlation is obtained if the following intermediates are assumed to occur on FMN reduction in 0.05 M H2SO4: FMNH., (FMN·FMNH2) and (FMNH.·FMNH2); it must also be assumed that the last named intermediate is non-reducible.  相似文献   

19.
The synthesis, IR and Raman spectroscopic study, and X-ray diffraction analysis of [Ni(HL1)3](L2)2 · 4H2O (I), where HL1 is phenylacetic acid hydrazide and L2 is the benzoate monoanion, have been performed. The structural units of a crystal of complex I are complex [Ni(HL1)3]2+ cations, (L2) anions, and crystallization water molecules. The nickel atom is coordinated to the three oxygen atoms at octahedron apices and the three nitrogen atoms of three bidentate chelate (О, N) ligands HL1 in cis,trans-meredianal (fac) conformation. The structural units of a crystal of complex I are bonded by a branched network of О–Н···О and N–H···O hydrogen bonds.  相似文献   

20.
The distribution of nitric acid between an aqueous phase of constant or variable ionic strength and a benzene solution of diphosphine dioxide can be explained by the following reactions H+a+ NO3-a+ DiPO0 ? D1PO·HNO30 H+a+ NO3-a+ DiPO·HNO30 ? DiPO·2 HNO30 At constant ionic strength, the stability constants K1″ were determined for the complexes 1,1-DiPO·HNO3 (98 ± 01 (M)-1), 1,4-DiPO·HNO3(44±3 (M)-1) and 1,5-DiPO·HNO3 (51 ± 1 (M)-1). The constants K11″ for the complexes 1,1-DiPO·2 HNO3 and 1,5-DiPO.2 HNO3 are respectively 035±001 (M)-1 and 62 ±0.05 (M)-1 at 25°. With an aqueous phase of variable ionic strength, values of K1'=54±7 (M)-2 for 1,5-D1PO.HNO3 and KII'=65 ± 04 (M)-2 for 1,5-DiPO·2 HN03 were obtained  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