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.
The kinetics of 2,2′-azobisisobutyronitrile (AIBN) initiated polymerization of styrene in N,N-dimethylformamide (DMF) at 60°C were investigated in the presence of dibromo(N,N-dimethylformamide)copper(II) complex. The complex was prepared in situ by mixing tetrakis (N,N-dimethylformamide)copper(II) perchlorate with LiBr in the molar ratio of 1 : 2. The equilibrium constant for [Cu(DMF)4]2+ + 2Br? ? Cu(DMF)2Br2 + 2DMF was calculated by the limiting logarithmic method as 1.80 × 103 L2 mol?2. The velocity constant at 60°C for the interaction of polystyryl radical with Cu(DMF)2Br2 is 7.46 × 104 L mol?1 s?1.  相似文献   

2.
The kinetics of charge-transfer (CT) polymerization of methyl methacrylate (MMA) in the presence of imidazole (Imy) and CCl4 was studied in dimethyl sulfoxide (DMSO) at 60°C. The rate of polymerization (Rp) is sensitive to the [CCl4] up to a concentration of 0.60 mol L−1, but at a higher concentration, it is practically independent of the [CCl4]. When [CCl4] > [Imy], Rp is proportional, to [MMA]1.45±0.15 and [Imy]0.53±0.04 and the average rate constant for the polymerization of MMA is 3.25±0.41 × 10−6 L mol−1 s−1. This article also reports the polymerization of MMA initiated by Imy and CCl4 and accelerated by hexakis (dimethyl sulfoxide) iron (III) perchlorate, [Fe(DMSO)6] (CIO4)3 (A), at 60°C. The presence of Fe(Imy)3+3 in the polymerization system produced well-defined induction periods. The rate constant at 60°C for the interaction of the poly(MMA) radical toward Fe(Imy)3+3 is 7.19 × 104 L mol−1 s−1. A probable reaction mechanism for the polymerization system has been postulated to explain the observed results. © 1996 John Wiley & Sons, Inc.  相似文献   

3.
The reaction of Fe(CO)5 with CF3COOH/(CF3CO)2O, or with CF3COOH, in DMF at about 140 °C produces the colourless compound 1, [Fe(DMF)6][Fe33-O) (CF3COO)6(DMF)3]2, constituted by the [Fe(DMF)6]2+ cation and by two [Fe33-O)(μ2-CF3COO)6(DMF)3] anions, with a triangular disposition of the three iron atoms, the central substantially coplanar oxygen atom being formulated as a singly charged anion O(–I) stabilized by complexation to iron(II).  相似文献   

4.
《Electrochimica acta》1986,31(2):185-192
Visible absorption spectra and the molar conductance curve of CoBr2 dissolved in N,N-dimethylformamide (DMF) have been determined at 25°C. Absorption spectra of ternary systems Co(ClO4)2 + Et4NBr + DMF and CoBr2 + chlorobenzene + DMF have also been determined at 25°C. The results indicate the formation of the CoBr(DMF)+5. CoBr3DMF complex electrolyte controlling the electrolytic properties of the solution. The stability constants of the individual bromo-complexes of cobalt(II) have been calculated.  相似文献   

5.
The water sorption characteristics of poly(ethylene terephthalate) (PET) amorphous samples of 250 μm thickness have been studied at various temperatures in a saturated atmosphere. Concerning diffusivity, one can distinguish the following two domains characterized by distinct values of the activation energy: ED ≈ 36 kJ mol−1 at T > 100°C, and ED ≈ 42 kJ mol−1 at T < 60°C, with a relatively wide (60–100°C) intermediary domain linked to the glass transition of the polymer. The crystallization of this latter occurs in the time scale of diffusion above 80°C but doesn't change the Fickian character of sorption curves. The equilibrium concentration m is an increasing function of temperature, but the solubility coefficient S decreases sharply with this latter, with the apparent enthalpy of dissolution ΔHs being of the order of −28 kJ mol−1 at T < 80°C and −45 kJ mol−1 at T > 80°C. Density measurements in the wet and dry states suggest that water is almost entirely dissolved in the amorphous matrix at T < 80°C but forms partially a separated phase at T > 80°C. Microvoiding can be attributed to crystallization-induced demixing. © 1999 John Wiley & Sons, Inc. J Appl Polym Sci 73: 1131–1137, 1999  相似文献   

6.
Polymerization of methyl methacrylate (MMA) by the charge-transfer complex formed by the interaction of 2,2′-bipyridine (bpy), MMA, and carbon tetrachloride (CCl4) was studied in dimethylsulfoxide (DMSO) at 60°C. The rate of polymerization (Rp) is sensitive to the [CCl4] at low concentration of CCl4, but at a higher concentration it is practically independent of [CCl4]. Rp is proportional to [MMA]1.45±0.04 and [bpy]0.52±0.04 when [CCl4] > [bpy], and the average rate constant, k, at 60°C for the polymerization of MMA was 7.14 ± 0.40 × 10−6 L mol−1s−1. Kinetic studies showed that the polymerization proceeds through free radical intermediates. This article also reports the polymerization of MMA initiated by bpy and CCl4 and accelerated by Lewis acid, hexakis (dimethylsulfoxide)iron(III) perchlorate [Fe(DMSO)6](ClO4)3 at 60°C. The glass transition temperature and molecular weights of the samples were investigated by using differential scanning calorimetry and gel permeation chromatography techniques, respectively. Probable reaction mechanisms are proposed to explain the observed results. © 1997 John Wiley & Sons, Inc. J Appl Polym Sci 64: 2097–2103, 1997  相似文献   

7.
This work reports the kinetics of anionic ring-opening polymerization of hexaethylcyclotrisiloxane (D3Et) with potassium trimethylsilanolate as initiator and diglyme and N,N-dimethylformamide (DMF) as promoters. The polymerization rate of D3Et is influenced by the nature of promoters and the reaction temperatures. With the use of DMF as a promoter, the polymerization activation energy is 107.89 kJ/mol, and the polymerization rate constant at 110°C is 0.08466 min−1 with [P]/[I] = 3.0. Gel permeation chromatography and 29Si NMR spectra showed that the intermolecular redistribution occurred during the late stage of polymerization, which facilitated the synthesis of high-molecular-weight polydiethylsiloxane (PDES). Differential scanning calorimetry analysis showed that PDES exhibited a glass transition temperature of −142°C and complex crystallization phenomena. Thermogravimetric analysis illustrated that the PDES that was prepared using this method had good thermal stability with onset decomposition temperatures of 483 and 452°C under nitrogen and air conditions, respectively. This work presents innovative approaches for achieving a more energy-efficient synthesis of PDES to meet the demands of industrial-scale production.  相似文献   

8.
《Electrochimica acta》1985,30(5):603-612
Visible absorption spectra and the molar conductance curve at NiCl2 dissolved in dimethylformamide (DMF) have been determined at 25°C. Absorption spectra of ternary system Ni(ClO4)2 + Et4NCl + DMF and NiCl2 + chlorobenzene + DMF have also been determined at 25°C. It has been found that at lower concentrations of NiCl2 monochloride nickel complex NiCl(DMF)+5 and outer-sphere ion-pair {NiCl(DMF)+5Cl}0 coexist in the solution. The presence of pseudotetrahedral complex NiCl3DMF and the existence of coordinative disproportionation equilibrium have been established in the more concentrated solutions of NiCl2. The formation constant of outer-sphere ion-pair equal to (1.33 ± 0.2) × 102 has been derived from the conductometric data. The equilibrium constant of disproportionation reaction equal to (3.89 ± 0.2) × 102 has been determined on the basis of the absorption spectra of NiCl2 in DMF.  相似文献   

9.
Radical copolymerization of zinc acrylate (ZnA2) with acrylonitrile (AN), initiated by As2S3–styrene complex(I), in dimethyl Sulphoxide (DMSO) at 90 ± 0.1°C for 1.0h under inert atmosphere, yields non-alternating copolymers. The kinetic expression is Rp ∝ [I]0.33 [ZnA2]0.25 [AN]0.44, i.e. the system follows non-ideal kinetics, which is due to primary radical termination as well as degradative chain transfer reactions. The values for activation energy (E) and k2p/kt are 128kJ mol−1 and 8.57 × 10−7 litre mol−1 s−1, respectively. Thermal stability, solubility in different solvents, and IR and NMR spectra have been evaluated.  相似文献   

10.
The solubility of potassium ferrate (K2FeO4) was measured in aqueous solutions of NaOH and KOH of total concentration 12 M containing various molar ratios of KOH:NaOH in the range 12:0 to 3:9. Several analytical methods were tested for the determination of ferrate concentration. The final method chosen consisted of potentiometric titration of the ferrate sample with an alkaline solution of As2O3. The assumption was made that ferrate dissociates in concentrated KOH solutions predominantly to KFeO4. The solubility constant, S, defined as the product of the molar concentration of the potassium ion, K+, and the ferrate anion, KFeO4, was found to be 0·044 ± 0·006 mol2 dm−6 for 20°C, 0·093 ± 0·004 mol2 dm−6 for 40°C and 0·15 ± 0·09 mol2 dm−6 for 60°C. From these results the heat of dissolution of K2FeO4 was calculated as −14·3 kJ mol−1. At 60°C the enhanced decomposition of the ferrate at the higher temperature led to a greater deviation in solubility values compared with data for either 20°C or 40°C.  相似文献   

11.
Chemical modification of polyvinyl chloride by nucleophiles is a versatile method for preparation of new functional polymers. Modification of PVC was carried out using different nucleophiles (Nu) in ethylene glycol (EG)/N,N-dimethylformamide (DMF) (1:1 by volume) solution. OH?, N3?, and SCN? were examined as nucleophiles in this study. The FTIR spectrum confirmed substitution of Cl by nucleophiles and also the elimination of HCl in the modification reaction of PVC. The glass transition temperatures (Tgs) of modified-PVC samples were in the order N3? > SCN? > OH? and were equal to 87.6, 86.8 and 78.15 °C, respectively. The T5% (the temperature at which weight loss is 5%) of PVC was 266 °C, while after 1 h modification by OH? it reduced to 197 °C. After alkaline modification, the scanning electron microscopy (SEM) images not only showed an increase in surface roughness and porosity, but also revealed a relatively large drop in average particle size from 160 to 80 µm. Molecular weight and molecular weight distribution were determined by gel permeation chromatography (GPC). The GPC results showed that the number average molecular weight (Mn) and weight average molecular weight (Mw) were decreased from 79,950 and 176,360 g mol?1 in crude PVC to 45,370 and 99,930 g mol?1, respectively, after 1 h modification by OH?. Alkaline treatment also decreased the mechanical strength of PVC.  相似文献   

12.
The coil–globule transition for poly(methyl methacrylate) (PMMA) has been studied in a theta solvent, acetonitrile (Θ = 45 °C). The viscosity of PMMA was measured as a function of temperature in the range 26–55 °C. The contraction and expansion of the molecular chains are determined using the measured viscosity values. The temperature dependence of the intrinsic viscosity can be represented by a master curve in a versus |τ|M w1/2 (g1/2 mol−1/2) plot, where τ = |T − Θ|/T is the reduced temperature and Mw‐is the weight‐average molecular weight. A universal plot of reduced viscosity versus reduced blob parameter (N/Nc) shows the attainment of the collapsed state below the theta temperature. The dimensions of PMMA in acetonitrile (Mw = 3.15 × 106 g mol−1) decrease to 63 % at 26 °C of those in the unperturbed state. The results in this work are compared with those previously published. © 2000 Society of Chemical Industry  相似文献   

13.
Random copolymers of P(MMA-co-styrene) were synthesized via single electron transfer-living radical polymerization (SET LRP) at 25 °C in N,N-dimethylformamide (DMF) and benzene using CCl4 as initiator and Fe(0) wire/N,N,N′,N′-tetramethyl-1,2-ethanediamine (TMEDA)/hydrazine (NH2NH2) complexes as catalyst in the presence of air. Fe(0) wire-mediated single electron transfer-living radical copolymerization of MMA and styrene represented a robust and versatile technique to synthesize the well-defined copolymers. The copolymerization rate was faster in DMF than in benzene, as determined by the apparent rate constants. The results showed that the copolymerization followed first-order kinetics model in the presence of polar DMF and non-polar benzene. The molecular weights increased linearly with the increase of monomer conversion with a narrow polydispersity index when the conversion was beyond 25 %. The polarity and the quantity of solvent had significant effects on the polymerization, and the apparent rate constants were 1.28 × 10?4, 1.21 × 10?4, and 9.23 × 10?5 s?1 in the order of DMF amount, 5, 10, and 15 mL. The conversion increased from 29.3 to 48.5 % and the polydispersity index (PDI) changed from 1.24 to 1.21 with [CCl4]0/[TMEDA]0 molar ratio changing from 1:0.5 to 1:5. The chain extension experiment demonstrated that the copolymerization exhibited a living characteristic.  相似文献   

14.
Solution terpolymerization of styrene (Sty), acrylonitrile (AN) and copper acrylate (CuA) has been carried out in dimethylformamide at 90°C for 4 h using p-acetyl benzylidene triphenylarsonium ylide as radical initiator. 1H nuclear magnetic resonance (NMR), IR and elemental analysis have been used to characterized the terpolymer. Analysis of kinetic data indicates the following rate equation: The overall activation energy is 38 kJ mol−1. The composition of terpolymer calculated from NMR and elemental analysis has been used to evaluate reactivity ratios as r1(Sty) = 5 ± 2 and r2(AN + CuA) = 0.4 ± 0.02 employing the Finemann–Ross method, which confirms its random origin. The terpolymer was thermally stable up to 2007deg;C.  相似文献   

15.
5‐Aminotetrazolium nitrate was synthesized in high yield and characterized using Raman and multinuclear NMR spectroscopy (1H, 13C, 15N). The molecular structure of 5‐aminotetrazolium nitrate in the crystalline state was determined by X‐ray crystallography: monoclinic, P 21/c, a=1.05493(8) nm, b=0.34556(4) nm, c=1.4606(1) nm, β=90.548(9)°, V=0.53244(8) nm3, Z=4, ϱ=1.847 g cm−3, R1=0.034, wR2 (all data)=0.090. The thermal stability of 5‐aminotetrazolium nitrate was determined using differential scanning calorimetry; the compound decomposes at 167 °C. The enthalpy of combustion (ΔcombH) of 5‐aminotetrazolium nitrate ([CH4N5]+[NO3]) was determined experimentally using oxygen bomb calorimetry: ΔcombH([CH4N5]+[NO3])=−6020±200 kJ kg−1. The standard enthalpy of formation (ΔfH°) of [CH4N5]+[NO3] was obtained on the basis of quantum chemical computations at the electron‐correlated ab initio MP2 (second order Møller‐Plesset perturbation theory) level of theory using a correlation consistent double‐zeta basis set (cc‐pVTZ): ΔfH°([CH4N5]+[NO3](s))=+87 kJ mol−1=+586 kJ kg−1. The detonation velocity (D) and the detonation pressure (P) of 5‐aminotetrazolium nitrate were calculated using the empirical equations by Kamlet and Jacobs: D([CH4N5]+[NO3])=8.90 mm μs−1 and P([CH4N5]+[NO3])=35.7 GPa.  相似文献   

16.
The enthalpies of combustion (ΔcombH) of dinitrobiuret (DNB) and diaminotetrazolium nitrate (HDAT‐NO3) were determined experimentally using oxygen bomb calorimetry: ΔcombH(DNB)=5195±200 kJ kg−1, ΔcombH(HDAT‐NO3)=7900±300 kJ kg−1. The standard enthalpies of formation (ΔfH°) of DNB and HDAT‐NO3 were obtained on the basis of quantum chemical computations at the electron‐correlated ab initio MP2 (second order Møller‐Plesset perturbation theory) level of theory using a correlation consistent double‐zeta basis set (cc‐pVTZ): ΔfH°(DNB)=−353 kJ mol−1, −1 829 kJ kg−1; ΔfH°(HDAT‐NO3)=+254 kJ mol−1, +1 558 kJ kg−1. The detonation velocities (D) and detonation pressures (P) of DNB and HDAT‐NO3 were calculated using the empirical equations by Kamlet and Jacobs: D(DNB)=8.66 mm μs−1, P(DNB)=33.9 GPa, D(HDAT‐NO3)=8.77 mm μs−1, P(HDAT‐NO3)=33.3 GPa.  相似文献   

17.
A novel siloxane-containing diamine, bis(p-aminophenoxy)methylphenylsilane (BAMPS), was synthesized from the condensation of dichloromethylphenylsilane with p-aminophenol in the presence of triethylamine. A series of BAMPS-based aromatic polyimides were prepared from BAMPS and various aromatic tetracarboxylic dianhydrides by the usual two-step procedure including ring-opening polyaddition to poly(amic acid)s and subsequent cyclodehydration to polyimides. The inherent viscosities of poly(amic acid)s IIIa - IIIf ranged from 0.09 to 0.36 dL g−1 in N,N-dimethylacetamide at a concentration of 0.5 g dL−1 at 30°C. The inherent viscosities of polyimides were between 0.06 and 0.32 dL g−1 in various solvents at 30°C. Polyimides, especially IVc and IVf , were soluble in a wide range of organic solvents such as N-methyl-2-pyrrolidinone, concentrated H2SO4, N,N-dimethylacetamide, N,N-dimethylformamide, and dimethyl sulfoxide. The polyimides were characterized by elementary analysis, IR spectra, TGA, and DSC. They also had glass transition temperatures ranging from 128 to 181°C. The 10% mass loss temperature was recorded in the range of 404–443°C in nitrogen and of 315–339°C in oxygen. © 1997 John Wiley & Sons, Inc.  相似文献   

18.
The isothermal kinetic of the release of nicotine from a poly(acrylic acid) (PAA) hydrogel was investigated at temperature range from 26°C to 45°C. Specific shape parameters of the kinetic curves, the period of linearity and saturation time were determined. The change in the specific shape parameters of the kinetic curves with temperature and the kinetic parameters of release of nicotine Ea and ln A were determined. By applying the “model fitting” method it was established that the kinetic model of release of nicotine from the PAA hydrogel was [1 − (1 − α)1/3] = kMt. The limiting stage of the kinetics release of nicotine was found to be the contracting volume of the interaction interface. The distribution function of the activation energy was determined and the most probable values of activation energies of 25.5 kJ mol−1 and 35 kJ mol−1 were obtained. Energetically heterogeneity of the interaction interface was explained by the existence of the two different modes of bonding the nicotine molecules onto the hydrogel network by hydrogen bond and electrostatic forces. © 2010 Wiley Periodicals, Inc. J Appl Polym Sci, 2011  相似文献   

19.
Reactions of H atoms with COS, CS2, and H2S were studied behind reflected shock waves at temperatures between 1170 K and 1830 K and pressures around 1.0 bar by applying atomic resonance absorption spectroscopy (ARAS) for time-resolved measurements of H atoms at Lα. The thermal decomposition of a few ppm ethyl iodide (C2H5I) was used as a H-atom source. In the presence of a large excess of the molecular reactant COS, CS2, or H2S, a consumption of H was observed which follows a pseudo first-order rate law. Rate coefficients for the reactions: were determined to be: k1 = 2.4 × 1014exp(–3415 K/T) cm3mol−1s−1 k2 = 1.4 × 1015exp(–9250 K/T) cm3mol−1s−1 k3 = 2.5 × 1014exp(–2890 K/T) cm3mol−1s−1  相似文献   

20.
The kinetic investigation of Hg(II)-promoted reaction between [Fe(CN)6]4− and 2,2′-bipyridine (Bipy) has been performed in anionic sodium dodecyl sulfate (SDS) micellar medium by recording the surge in absorbance at 400 nm, corresponding to ultimate reaction product [Fe(CN)4 Bipy]2− using UV–visible spectrophotometer. Pseudo-first-order condition has been used to examine the progress of reaction as a function of temperature, [Fe(CN)64−], ionic strength, [SDS], pH, [Hg2+], and [Bipy] by changing one parameter at a time. The results exhibit that [Hg2+], [SDS], and pH are the decisive parameter showing maximum reaction rate at 1.5 × 10−4 mol dm−3, 6.0 × 10−3 mol dm−3, and 3.8, respectively. [Fe(CN)6]4− does not show any appreciable effect on the critical micellar concentration (CMC) of SDS as the polar head of SDS and [Fe(CN)6]4− both are negatively charged. Variable order kinetics was observed for [Fe(CN)6]4− and Bipy in their examined concentration range. The reverse response observed in the reaction rate with [KNO3] shows a negative salt effect. The K+ provided by K4[Fe(CN)6] and KNO3 decreases the repulsion between the negatively charged heads of the surfactant molecules thereby decreasing the CMC of SDS. The negative value for the entropy of activation also supports the interchange dissociative (Id) mechanism recommended by us.  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