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.
In this paper, the mass transfer coefficients for trichloroethylene (TCE), toluene (TOL) and dimethyl sulfide (DMS) are experimentally determined for different porous and composite membranes. For polypropylene/polyvinylidenedifluoride porous layer/thin film polydimethylsiloxane dense layer composite membranes, membrane mass transfer coefficients are 2.55E−03, 2.82E−03 and 2.90E−03 m/s for TCE, TOL and DMS in N2 at 30.0 ± 0.1 °C, respectively. For polyester/polyacrylonitrile porous layer/thin film polydimethylsiloxane dense layer composite membranes, they are higher, namely 4.28E−03, 4.55E−03 and 4.81E−03 m/s for TCE, TOL and DMS in N2 at 30.0 ± 0.1 °C, respectively. Analysis of the contribution of the dense layer of both composite membranes to the total membrane resistance for mass transfer, showed that this contribution was small for both composite membranes. The higher mass transfer coefficients of the thin film polydimethylsiloxane composite membranes from this study in comparison to others from the literature are primarily due to improvement of the mass transfer characteristics of the porous layer. Analysis of the mass transfer characteristics of the different porous layers of which the total porous layer is composed, showed that the contribution of the porous “backing” layer for mechanical support can be substantial in comparison to the porous layer in contact with the dense layer.  相似文献   

2.
The 1H NMR technique was applied for the measurement of the isomerization rates of N-ethyl-N-[(benzotriazol-1-yl)methyl]aniline ( 4 ) and 4-butyl-N-[(benzotriazol-1-yl)methyl]aniline ( 7 ) to the corresponding benzotriazol-2-yl isomers in dioxane-d8 at 35°C. The rate constants obtained for pure dioxane-d8 were 1.62 and 0.28 h?1 for 4 and 7 , respectively. For both compounds, addition to acetic acid to the dioxane solutions accelerated the isomerizations whereas addition of triethylamine retarded it strongly. Addition of water slowed the isomerization of 4 but accelerated that of 7 : the different effects operating in the two cases are discussed and rationalized. © 1995 John Wiley & Sons, Inc.  相似文献   

3.
Aqueous solutions of N,Nʹ-diethylethanolamine (DEEA) are prospective high-capacity CO2-capturing solvents. Their reactivity can be enhanced by promotion with absorption activators. Two polyamines were chosen as activators in this work, viz. (methylamino)propylamine (MAPA) and diethylene triamine (DETA). In a stirred cell reactor, kinetics of CO2 absorption into aqueous DEEA/MAPA and DEEA/DETA mixtures was studied at 303 K. The molarity of DEEA was varied in the 2–2.5 M range, whereas the polyamine concentration was changed between 0.1 and 0.5 M. Pseudo–first-order rate constants were reported. Second-order rate constants for the CO2 reactions with MAPA and DETA were determined too. DETA reacts faster than MAPA.  相似文献   

4.
The rates of heat release in the nitrogen dioxide—n-decane system at a molar ratio of nitrogen oxides ton-decane (β) from 2.4·10−3 to 3.1 and gaseous volumes per mole ofn-decane (V(g)) equal to 0.05–4.5 were studied in the 55.2–92.8 °C temperature range. The initial rate of the process is determined by the interaction of NO2 withn-decane. The equilibrium constants of dissociation of N2O4 inn-decane and Henry's constants of NO2 and N2O4 in ann-decane solution were determined by complex analysis of the thermodynamic equilibrium in the NO2n-decane system and dependences of the initial rates onV(g) and β. The experimentally observed self-acceleration of the process in the region of high β and lowT values was suggested to be due to the reaction of N2O4 with intermediate oxidation products. The rate constants of the reaction of NO2 withn-decane were compared with analogous values determined in its mixtures with HNO3 solutions. Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 10, pp. 1789–1794, October, 1997.  相似文献   

5.
The standard partial molar entropy of the aqueous tetrabutylammonium cation, not known previously, has now been obtained, based on the molar entropy of two of its crystalline salts, the iodide and the tetraphenylborate, recently determined experimentally for this purpose. The calculation required also published molar enthalpies of solution and solubilities of these two salts as well as of the perchlorate. The choice of the anions depended mainly on the limited solubilities of the examined salts in water, facilitating the estimation of the relevant activity coefficients. The result is S(Bu4N+, aq) = (380 ± 20) J · K−1 · mol−1 at T = 298.15 K, on the mol · dm−3 scale and based on S(H+, aq) = (−22.2 ± 1.2) J · K−1 · mol−1 (yielding the ‘absolute’ value). The molar entropy of this cation in the ideal gas standard state, S(Bu4N+, g) = (798 ± 8) J · K−1 · mol−1 then yielded the molar entropy of hydration ΔhydS (Bu4N+) = (−418 ± 23) J · K−1 · mol−1.  相似文献   

6.
Described in this communication is a new synthesis of N-methyl-4-(vinylaryl)quinolinium salts by cyclization of the aminoketones, Ar-CH=CH-CO-CH2-CH2-N(CH3)-C6H4-X. The aminoketones were obtained by replacement of the alkyiamino group of the ketoarylidene Mannich bases with various N-methylarylamines. The U.V. spectra of the compounds were obtained. The photosensitivity characteristics in solution are reported. These data show the possibility of dimerization and probable transcis isomerization on exposure to visible and ultraviolet light.  相似文献   

7.
The crystal structures of the isostructural compounds α-CF3HgN3 (I) and CF3HgNCO (II) were determined by X-ray methods. Centrosymmetric dimers are formed with weak Hg---n′ bonds of 2.74(2) Å in I and 2.88(2) Å in II. The Hg---N distance and Hg---N---N angle are 2.02(2) Å and 127(2)° in the monomeric fragment of I, the corresponding values in II are 2.03(2) Å and 137(2)°. Raman and, in part, IR spectra of both compounds in the solid state were recorded and assigned for Cs and C2h symmetry of the monomers and dimers, respectively. A normal coordinate analysis proved that the stretching vibrations of the (HgN)2 four-membered ring appeared at ≈400 and ≈75 cm-1; the force constants amounted to ≈2.0 for the short and to ≈0.25 X 102 N/m for the long HgN bonds.  相似文献   

8.
The α-tocopheroxyl radical was generated voltammetrically by one-electron oxidation of the α-tocopherol anion (r1/2=−0.73 V versus Ag|Ag+) that was prepared by reacting α-tocopherol with Et4NOH in acetonitrile (with Bu4NPF6 as the supporting electrolyte). Cyclic voltammograms recorded at variable scan rates (0.05–10 V s−1), temperatures (−20 to 20°C) and concentrations (0.5–10 mM) were modelled using digital simulation techniques to determine the rate of bimolecular self-reaction of α-tocopheroxyl radicals. The k values were calculated to be 3×103 l mol−1 s−1 at 20°C, 2×103 l mol−1 s−1 at 0°C and 1.2×103 l mol−1 s−1 at −20°C. In situ electrochemical-EPR experiments performed at a channel electrode confirmed the existence of the α-tocopheroxyl radical.  相似文献   

9.
In order to elucidate the photo-decomposition mechanism of polyurethane based on polyester diol-diphenylmethane-p,p′-diisocyanate, the effects of triplet quenchers, piperylene and oxygen on the photo-decomposition of the polymer, methylene bis (ethyl N-phenylcarbamate) (MEPC) and ethyl N-phenylcarbamate (EPC) were examined in solution. Energy levels and lifetimes of the excited states of these compounds were also determined.Piperylene and oxygen did not affect the photo-decomposition of the samples examined. The results imply that the photo-decomposition of the polymer starts from the excited singlet state. The energy levels and lifetimes for the photo-decomposition of the polymer were as follows: the excited singlet state (S1): 98·6 kcal/mol (3·2 nsec): the excited triplet state (T1): 76·7 kcal/mole (2·9 sec).  相似文献   

10.
Kinetics and mechanism of the reactions of methyl diazoacetate, dimethyl diazomalonate, 4-nitrophenyldiazomethane, and diphenyldiazomethane with sulfonium ylides and enamines were investigated by UV-Vis and NMR spectroscopy. Ordinary alkenes undergo 1,3-dipolar cycloadditions with these diazo compounds. In contrast, sulfonium ylides and enamines attack at the terminal nitrogen of the diazo alkanes to give zwitterions, which undergo various subsequent reactions. As only one new bond is formed in the rate-determining step of these reactions, the correlation lg k2(20 °C)=sN(N+E) could be used to determine the one-bond electrophilicities E of the diazo compounds from the measured second-order rate constants and the known reactivity indices N and sN of the sulfonium ylides and enamines. The resulting electrophilicity parameters (−21<E<−18), which are 11–14 orders of magnitude smaller than that of the benzenediazonium ion, are used to define the scope of one-bond nucleophiles which may react with these diazoalkanes.  相似文献   

11.
A new series of mono- and di-substituted ruthenium–polypyridine complexes of the type cis-[RuIII,II (bpy)2(L1)(L2)] n + (L1 = Cl, bta or py; L2 = bta; bpy = 2,2-bipyridine; bta = benzotriazole; py = pyridine) has been prepared, isolated as hexafluorophosphate salts, and investigated in organic solutions by means of cyclic voltammetry and spectroelectrochemistry. The chemical oxidation of all the benzotriazole derivatives, starting from cis-[RuII(bpy)2(L1)(bta)] x (x = +1 or +2), leads essentially to N(3)-bound products, i.e., cis-[RuIII(bpy)2 (L1)(N(3){bta})] x+1 isomers. Nevertheless, while the benzotriazole-monosubstituted species undergoes an intramolecular isomerization, N(3) N(2), accompanying the electrochemical reduction centered on the metal ion, RuIII RuII, the disubstituted derivatives do not display any spectral or electrochemical evidence of linkage isomerism. The equilibrium and kinetic constants for the isomerization were determined from the cyclic voltammograms at several scan rates, according to an electrochemical–chemical (EC) coupling scheme. The data were compared with a set of constants and parameters obtained previously for a series of ruthenium and iron complexes. The experimental results were found to be quite consistent with theoretical calculations, and reflect the importance of -backbonding interactions in the stabilization of the metal-centered reduced state (MII) on such species with low-spin d6 configuration.  相似文献   

12.
The kinetics of the oxidation of 4,6-dimethyl-2-mercaptopyrimidine (DMP) by Ag(cyclam)2+ were studied in buffer solutions from pH 5.8 to 7.2 at constant ionic strength of 0.10?M?(NaClO4). The reaction is observed to be first-order with respect to [Ag(cyclam)2+] and to [DMP]. However, the reaction rate is affected by the pH of the solution owing to the acid–base equilibrium of the thiol. The mechanism postulated to account for the kinetics includes an acid–base equilibrium and oxidation of thiol (RSH) and thiolate ion (RS?) by Ag(cyclam)2+ to RS· radicals which undergo rapid dimerization to form disulfide (RSSR). From the postulated mechanism and the observed kinetics a rate expression was derived, and second-order rate constants and activation parameters were calculated. The pK a values of the acid dissociation reaction of DMP were also determined at four temperatures using spectrophotometric methods, and thermodynamic parameters calculated from the K a values.  相似文献   

13.
The equilibrium concentrations of all reaction products emerging from the hydrolysis ofN-bromo compounds in the presence of bromide and thereby also the hydrolysis constants (K 1) have been calculated from the absorbance at 392.8 nm, thepH-value and the initial concentrations of theN-bromo compound and the bromide. The following compounds have been investigated:N-bromo-succinimide:K 1=2.2·10–6, 1,3-dibromo-5,5-dimethylhydantoin:K 1=1.7·10–5,N-bromoacetamide:K 1=1.8·10–6,N-bromo-monochloroacetamide: 5.2·10–6,N-bromo-dichloroacetamide:K 1=8.9·10–6 andN-bromo-trichloroacetamide:K 1=1.8·10–5. The precision of the method, which is mainly suited for weak hydrolizingN-bromocompounds (K 1<10–4) are discussed and the overall error of the calculated values was found to be in the range of ±5–12%. The reactivities in aqueous solution of the most frequently usedN-bromo compounds are compared by means of the calculated HOBr equilibrium concentrations. The differences to be expected on the basis of the latters are at concentrations >10–5 mol/l rather great, while they can be neglected in very dilute solutions (-10–6 mol/l).
  相似文献   

14.
A new procedure to measure the equilibrium constants for the dimerization (homochiral and heterochiral) reactions of enantiomers in solution was applied to two different compounds, namely omeprazole and Pirkle's alcohol, both in CHCl3. This procedure is based on the measurement of the optical rotation of the solution as a function of its composition, which exhibits a nonlinear dependence on the enantiomer enrichment. Such nonlinearity depends on the extent of dimerization, and it is a strong effect in the case of omeprazole, whose equilibrium constants are 14.0±3.5 l/mol and 25.2±4.0 l/mol for the formation of homochiral and of heterochiral dimers, respectively. Pirkle's alcohol exhibits a weaker effect that allows only to estimate the order of magnitude of these constants, i.e., ca. 0.10 and 0.34 l/mol, respectively.  相似文献   

15.
TiCl4‐induced Baylis–Hillman reactions of α,β‐unsaturated carbonyl compounds with aldehydes yield the (Z)‐2‐(chloromethyl)vinyl carbonyl compounds 5 , which react with 1,4‐diazabicyclo[2.2.2]octane (DABCO), quinuclidine, and pyridines to give the allylammonium ions 6 . Their combination with less than one equivalent of the potassium salts of stabilized carbanions (e.g. malonate) yields methylene derivatives 8 under kinetically controlled conditions (SN2’ reactions). When more than one equivalent of the carbanions is used, a second SN2’ reaction converts 8 into their thermodynamically more stable allyl isomers 9 . The second‐order rate constants for the reactions of 6 with carbanions have been determined photometrically in DMSO. With these rate constants and the previously reported nucleophile‐specific parameters N and s for the stabilized carbanions, the correlation log k (20 °C)=s(N + E) allowed us to calculate the electrophilicity parameters E for the allylammonium ions 6 (?19<E <?18). The kinetic data indicate the SN2’ reactions to proceed via an addition–elimination mechanism with a rate‐determining addition step.  相似文献   

16.
The phase relations in the system In2O3–TiO2–MgO at 1100 and 1350°C are determined by a classical quenching method. In this system, there are four pseudobinary compounds, In2TiO5, MgTi2O5 (pseudobrookite type), MgTiO3 (ilmenite type), and Mg2TiO4 (spinel type) at 1100°C. At 1350°C, in addition to these compounds there exist a spinel-type solid solution Mg2−xIn2xTi1−xO4 (0≤x≤1) and a compound In6Ti6MgO22 with lattice constants a=5.9236(7) Å, b=3.3862(4) Å, c=6.3609(7) Å, β=108.15(1)°, and q=0.369, which is isostructural with the monoclinic In3Ti2FeO10 in the system In2O3–TiO2–MgO. The relation between the lattice constants of the spinel phase and the composition nearly satisfies Vegard's law. In6Ti6MgO22 extends a solid solution range to In20Ti17Mg3O67 with lattice constants of a=5.9230(5) Å, b=3.3823(3) Å, c=6.3698(6) Å, β=108.10(5)°, and q=0.360. The distributions of constituent cations in the solid solutions are discussed in terms of their ionic radius and site preference effect.  相似文献   

17.
The synthesis and characterisation of Cu(meclof)2H2O and Cu(meclof)2L2 (meclof = meclofenamate; L = 2-pyridylcarbinol (2-pyca), 3-pyridylcarbinol (3-pyca), nicotinamide (na), N,N-diethylnicotinamide (dena) are reported. The characterisation of the compounds were based on elemental analyses, electronic, IR and EPR spectra. The carboxyl group of the meclofenamate anions coordinates to the Cu(II) atom as an unidentate or as a chelating ligand. The crystal and molecular structures of one of the products, namely Cu(meclof)2(2-pyca)2 were measured. The EPR spectra of the studied complexes show they are monomeric, except for Cu(meclof)2 · H2O which shows triplet state feature. On the basis of the spectroscopic parameters observed, a monomeric structure with a tetragonally Jahn–Teller distorted octahedron around the Cu(II) atom is deduced for Cu(meclof)2L2 and for Cu(meclof)2 · H2O a dimeric structure is proposed. The degree of distortion in the series of the Cu(meclof)2L2 complexes increases in the order of L: na < 2-pyca < 3-pyca < dena.  相似文献   

18.
Zn1−xMgxO particles were prepared using zinc and magnesium oxalate precursor by co-precipitated method. The lattice constants of Zn1−xMgxO proved that the interstitial Mg formed at 500 °C and Mg replaced Zn in ZnO tetrahedral coordination at 800 °C. Compared with the ZnO, the absorbing band edge of the Zn1−xMgxO displayed blue shifts. The room temperature photoluminescence was similar to ZnO and variation of Mg content did not change the shape or peak position of the emission spectra markedly when it was annealed at 500 °C. However, its blue emission band disappeared, and a relatively strong green light emission at 498 nm appeared after annealed at 800 °C. The photoluminescence intensity ratios I(green)/I(UV) of Zn1−xMgxO varied with Mg content and the green light emission peak shifted from 498 nm to 472 nm when Mg content increased from 0 to 2.0 at.%.  相似文献   

19.
Adsorption (at a low temperature) of nitrogen on the protonic zeolite H-Y results in hydrogen bonding of the adsorbed N2 molecules with the zeolite Si(OH)Al Brønsted-acid groups. This hydrogen-bonding interaction leads to activation, in the infrared, of the fundamental N–N stretching mode, which appears at 2334 cm−1. From infrared spectra taken over a temperature range, the standard enthalpy of formation of the OH···N2 complex was found to be ΔH0 = −15.7(±1) kJ mol−1. Similarly, variable-temperature infrared spectroscopy was used to determine the standard enthalpy change involved in formation of H-bonded CO complexes for CO adsorbed on the zeolites H-ZSM-5 and H-FER; the corresponding values of ΔH0 were found to be −29.4(±1) and −28.4(±1) kJ mol−1, respectively. The whole set of results was analysed in the context of other relevant data available in the literature.  相似文献   

20.
Rate constants kiso of the thermal cis‐trans isomerization of four 4,4’‐nitro‐aminoazobenzenes with different amino groups have been determined in homogeneous aprotic solvents and polyglykol oligomers, primarily by means of conventional flash photolysis. The rate constants have been correlated with polarity (according to λmax from UV/Vis absorption spectra of the trans isomers) and bulk viscosity of the solvents. Qualitative conclusions about the influence of varying concentrations of water with respect to polarity and hydrogen bonding on kiso‐ and λmax‐values in acetone/water mixtures were derived. Based on these results the data from microheterogeneous solutions have been interpreted. In microheterogeneous water/surfactant solutions kiso‐values of selected azo dyes were strongly dependent on the concentrations of SDS, Triton®X‐100, C12EO8 in water, and varied with the composition of bicontinuous microemulsions of Igepal® CA‐520/ heptane/water. The large spread of isomerization rate constants is in part due to varying microviscosity. Replacement of H2O by D2O in aqueous surfactant solutions produced surprisingly large kinetic solvent isotope effects. © 1999 John Wiley & Sons, Inc. Int J Chem Kinet 31: 337–350, 1999  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