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.
Head-to-head (h-h) poly(acrylic acid) (PAA) and some h-h poly(alkyl acrylates) (PRA) with methyl, ethyl, n-propyl, n-butyl, isobutyl and 2-ethylhexyl substituents were prepared by hydrolysis or esterifications of the alternating copolymer of ethylene with maleic anhydride. In general, these esterification reactions became increasingly difficult as the carbon chain in the alcohols lengthened or branched. The softening, glass transition, and degradation temperatures of the h-h polymers obtained were somewhat higher than those of the corresponding head-to-tail (h-t) polymers. The main degradation products of both h-h and h-t PRA were identified by pyrolytic gas chromatography as the alcohol and monomer. In addition, the relative ratios of the amounts of alcohol to monomer were larger for h-h than for the corresponding h-t polymers.  相似文献   

2.
The rate of the perchloric acid hydrolysis of aqueous ethyl and butyl vinyl ethers at 25.0°C, in the presence of micellar aggregates [anionic, sodium dodecyl sulfate (SDS); cationic, cetyl trymethyl ammonium bromide (CTAB); and nonionic, polyoxyethylen? 23? dodecanol, (Brij 35)], has been studied. Negligible effects were observed in the cases of cationic and nonionic micelles. Anionic micelles produce an enhancement in the reaction velocity, and the rate constants go through maxima with increasing SDS concentration. These maxima disappear in the presence of excess sodium perchlorate. All these facts are interpreted quantitatively by means of the pseudo-phase ion-exchange model.  相似文献   

3.
A new, highly sensitive and simple kinetic method for the determination of thyroxine was proposed. The method was based on the catalytic effect of thyroxine on the oxidation of As(III) by Mn(III) metaphosphate. The kinetics of the reaction was studied in the presence of orthophosphoric acid. The reaction rate was followed spectrophotometrically at 516 nm. It was established that orthophosphoric acid increased the reaction rate and that the extent of the non-catalytic reaction was extremely small. A kinetic equation was postulated and the apparent rate constant was calculated. The dependence of the reaction rate on temperature was investigated and the energy of activation and other kinetic parameters were determined.

Thyroxine was determined under the optimal experimental conditions in the range 7.0 × 10−9 to 3.0 × 10−8 mol L−1 with a relative standard deviation up to 6.7% and a detection limit of 2.7 × 10−9 mol L−1. In the presence of 0.08 mol L−1 chloride, the detection limit decreased to 6.6 × 10−10 mol L−1. The proposed method was applied for the determination of thyroxine in tablets. The accuracy of the method was evaluated by comparison with the HPLC method.  相似文献   


4.
4-(1,4,7,10-Tetraazacyclotetradec-1-yl)methylbenzoic acid (cycmba, 1) has been synthesized, as a step towards the eventual development of sequence-specific hydrolytic complexes. A cobalt(III) complex of 1, [Co(cycmba)Cl2]Cl.1.5H2O (.1.5H2O) was found to be active against both an activated phosphodiester compound, bis(nitrophenyl)phosphate (BNPP), and supercoiled DNA. The presence of the benzoate group depresses the rate of hydrolysis of the ligand-Co(III) system at neutral pH, as confirmed by the kinetics results of a methyl ester analog. The ability of (2.1.5H2O) to bind to solid substrates and remain active was also demonstrated by attachment of the molecule to agarose beads.  相似文献   

5.
This paper describes a highly sensitive, selective catalytic-kinetic-spectrophotometric method for the determination of copper(II) concentration as low as 6 ng ml−1. The method is based on the catalytic effect of copper(II) on the oxidation of citric acid by alkaline hexacyanoferrate(III). The reaction was followed by measuring the decrease in absorbance of hexacyanoferrate(III) at 420 nm (λmax of [Fe(CN)6]3−,  = 1020 dm3 mol−1 cm−1). The dependence of rate of the indicator reaction on the reaction variables has been studied and discussed to propose a suitable mechanism to get a relation between the reaction rate and [Cu2+]. A fixed time procedure has been used to obtain a linear calibration curve between the initial rate and lower [Cu2+] or log[Cu2+] in the range 1 × 10−7 to 4 × 10−4 mol l−1 (6.35-25,400 ng ml−1). The detection limit has been calculated to be 4 ng ml−1. The maximum average error is 3.5%. The effect of the presence of various cations, commonly associated with copper(II) and some anions has also been investigated and discussed. The proposed method is sensitive, accurate, rapid and inexpensive compared to other techniques available for determination of copper(II) in such a large range of concentration. The new method has been successfully applied for the determination of copper(II) in various samples.  相似文献   

6.
The self-organization of systems based on the nonionic surfactant Triton X-100, polyethyleneimine, and lanthanum ions has been investigated. The micellar solutions of Triton-X-100 alone and the Triton-X-100-polyethyleneimine binary systems inhibit the hydrolysis of phosphonic acid esters. The Triton-X-100-polyethyleneimine-La(III) ternary system with a certain component ratio exerts a considerable catalytic effect, raising the rate of the reaction by more than three orders of magnitude relative to the rate of the alkaline hydrolysis of the phosphonates.  相似文献   

7.
The water-soluble monomers, 1-methyl-4-vinylimidazole, 1-methyl-5-vinylimidazole, 1-ethyl-5-vinylimidazole, and 1-propyl-5-vinylimidazole have been synthesized, polymerized, and copolymerized with 4(5)-vinylimidazole. The copolymers were characterized by 14C-labeling, NMR, pKa determination and viscosity measurements. The monomer reactivity ratios determined by 14C counting are r1 = 1.04; r2 = 0.94 [M1 = 4(5)-vinylimidazole, M2 = 1-methyl-4-vinylimidazole] and r1 = 1.01; r2 = 0.86 [M1 = 4(5)-vinylimidazole, M2 = 1-methyl-5-vinylimidazole]. The esterolytic activity of the copolymers for the hydrolysis of p-nitrophenyl acetate (PNPA) at pH 7–8 in 28.5% ethanol–water was higher than that of the mixtures of homopolymers. At pH 5–6 the esterolytic activities of the copolymers and the mixtures were similar. The most efficient esterolytic activity for PNPA hydrolysis at pH 7.11 in 28.5% ethanol–water occurred for copolymers containing 75 mole % 4(5)-vinylimidazole and for copolymers containing 1-methyl-4-vinylimidazole rather than 1-methyl-5-vinylimidazole.  相似文献   

8.
Catalytic activity of ferric oxide was evaluated in the oxidative dehydrogenation of n-butane. The apparent activation energy determined in the kinetic region was 37.7 kcal/mol. The activity of this oxide was undetectable up to 430°C. Reaction product distributions are shown as functions of temperature in the range of 430–650°C.
(III) -. 37,7 /. 430°C. 430–650°C.
  相似文献   

9.
The catalytic activity of acids in the chloromethylation of toluene has been studied in a complex mixture of toluene, paraformaldehyde, hydrochloric acid, acetic acid and zinc chloride.A relation between the Hammett acidity function and the initial reaction rate has been found: r0 = hmf(k,C) m = 1.25 where: h = 10H0; H0 being the Hammett acidity function.This relation describes the influence of three acids, acetic acid, hydrochloric acid and zinc chloride, on the initial reaction rate in the investigated range of variables CT0 = 0.075 M, CF0 = 2.5 M; 30% ⩽ CHAc ⩽ 50%, 5% ⩽ CZnCl2 ⩽ 15% and 2 M ⩽ CHCl ⩽ 4 M.These results are useful in the evaluation of the reaction rate for the chloromethylation of toluene as a function of the proton activities. Finally, some mechanistic implications are discussed.  相似文献   

10.
The hydrolysis of hydroxymethyl cyanoguanidine and methoxymethyl cyanoguianidine in acidic aqueous media were studied by analyzing the eluted formaldehyde and the reaction products. Dihydroxymethyl cyanoguanidine released two formaldehydes with different rates because one of the two hydroxymethyl groups is intramolecularly interactive through the hydrogen bond. Methanol dissociation from methoxymethyl cyanoguanidine was seven to eight times faster than the dissociation of hydroxymethyl group, and the overall hydrolysis kinetics were similar to those of hydroxymethyl cyanoguanidine.  相似文献   

11.
The branching reaction in the radical polymerization of vinyl acetate was studied kinetically. Branching occurs by polymer transfer as well as terminal double-bond copolymerization. The chain-transfer constants to the main chain (Cp,2) and to the acetoxy methyl group (Cp,1) on the polymer were calculated on the basis of the experimental data described in the preceding paper giving Cp,2 = 3.03 × 10?4, Cp,1 = 1.27 × 10?4 at 60°C, and Cp,2 = 2.48 × 10?4, Cp,1 = 0.52 × 10?4 at 0°C. Chain transfer to monomer is important with respect to the formation of the terminal double bond. The total values of transfer constants to the α- or β-position in the vinyl group and the acetoxymethyl group in vinyl acetate was determined to be 2.15 × 10?4 at 60°C. The transfer constant to the acetyl group in the monomer (Cm,1) was also evaluated to be 2.26 × 10?4 at 60°C from the quantitative determination of the carboxyl terminals in PVA. These facts suggest that the chain-transfer constant to the α- or β-position in the monomer (Cm,2) is nearly equal to zero within experimental error. Copolymerization reactivity parameters of the terminal double bond were also estimated. In conclusion, it has become clear that the formation of nonhydrolyzable branching by the terminal double-bond reaction can be almost neglected, and hence that the long branching in PVA is formed only by the polymer transfer mechanism. On the other hand, a large number of hydrolyzable branches in PVAc are prepared by the terminal double-bond reaction rather than by polymer transfer.  相似文献   

12.
A poly(methacrylamide-co-methylmethacrylate) (abbreviated PMAA-MMA) polymer support was studied for supporting a heteropolyacid (tungstophosphoric acid, H3PW12O40) with its surface positively charged in the polymerization step. PMAA-MMA supports could be obtained in a porous form by eliminating template reagent molecules (benzylmalonic acid) combined with properly selected monomer (methacrylamide). The amount of amine groups in PMAA-MMA directly determined the amount of H3PW12O40 impregnated, because the amine groups induced a positive charge on the PMAA-MMA surface. Finally, H3PW12O40/PMAA-MMA showed better acid catalytic activities than unsupported H3PW12O40 in alkylation of 1,3,5-trimethylbenzene with cyclohexene, which confirmed that PMAA-MMA supported H3PW12O40 effectively.  相似文献   

13.
RuCl3 can further catalyze the reaction between hexacyanoferrate(III) and iodide ions, which is already catalyzed by the hydrogen ions obtained from perchloric acid. Rate, when the reaction is catalyzed only by the hydrogen ions, was separated graphically from the rate when ruthenium(III) and H+ ions both catalyze the reaction. Reactions studied separately in the presence as well as in the absence of RuCl3 under similar conditions were found to follow second order kinetics w.r.t. [I]. While the rate showed direct proportionality w.r.t. [Fe(CN)6]3− and [RuCl3]. At low concentrations the reaction shows direct proportionality with respect to [H+] which tends to become proportional to the square of hydrogen ion concentrations. External addition of [Fe(CN)6]4− ions retards the reaction velocity while change in ionic strength of the medium has no effect on the rate. With the help of the intercept of the catalyst graph, extent of the reaction, which takes place without adding ruthenium(III) was calculated and it was in accordance with the values obtained from the separately studied reaction in which only H+ ions catalyze the reaction. It is proposed that ruthenium forms a complex, which slowly disproportionates into the rate-determining step. Arrhenius parameters at four different temperatures were also calculated.   相似文献   

14.
15.
A manganese(III) complex of tetraphenylporpholactone, Mn(TPPL)Cl, was synthesized and characterized, including by single-crystal X-ray diffraction; the catalytic activity of this complex for olefin epoxidation reactions is compared with that of manganese(III) tetraphenylporphyrin chloride, Mn(TPP)Cl.  相似文献   

16.
The kinetics of the acid hydrolysis of chromatopenta-amminecobalt(III) ion has been studied using a stopped-flow method over the acidity range 0.01≤[H+]<-1.0 mol dm−3 and 20.0°C<-ϕ<-30.0°C at ionic strengths 0.5 and 1.0 mol dm−3 (LiNO3). These studies reveal that the complex is first protonated and subsequently hydrolysed to the aquapentaammine cobalt(III) ion. The rate constants for the hydrolysis of the mono and diprotonated species at 25°C are 0.83±0.01 s−1 and (1.60±0.02)×104 mol−1 dm−3 s−1, respectively. TMC 2664  相似文献   

17.
Summary The decomposition of piperidinium hexathiocyanatochromate(III), (pipH)3[Cr(NCS)6](s), into Cr(NCS)3(s) and pipHSCN(g) has been studied isothermally and nonisothermally using t.g.a. Data from isothermal studies were analysed according to 17 different kinetic models and the (, T) data from nonisothermal experiments were analysed using 12 rate laws by the procedure of Reich and Stivala. It was found that while a first-order rate law gave the best fit to the data obtained from isothermal and nonisothermal experiments most frequently, considerable variation exists for both types of experiments. Using the first order model, the activation energy was found to be 77.2 ±4.4kJ mol–1.  相似文献   

18.
The kinetic study of the hydrolysis reaction of Z-substituted phenyl hydrogen maleates (Z = H, m-CH3, p-CH3, m-Cl, p-Cl and m-CN) was carried out in aqueous solution, and the results were complemented with theoretical studies. Under some experimental conditions, two kinetic processes were observed. One of them was ascribed to maleic anhydride formation and the other to the anhydride hydrolysis. The Br?nsted-type plot for the leaving-group dependence was linear with slope beta(lg) = -1. The experimental results are consistent with a mechanism that involves significant bond breaking in the rate-limiting transition state (alpha(lg) = 0.64). Theoretical results for the reaction in the gas phase showed an excellent Br?nsted-type dependence with a beta(lg) of -1.03. A tetrahedral intermediate (TI) could not be found through DFT gas-phase studies (B3LYP/6-311+G*). Calculations carried out within a continuous solvation model or with discrete water molecules failed to find a stable TI. With both models, a flat region on the potential-energy surface is found and a tight optimization of the structures led back to starting materials. The theoretical results do not discard the possible existence of an unstable intermediate on the free-energy surface, but the analysis of the whole body of results compared with other acyl transfer reactions lead us to suggest that an enforced concerted mechanism is the most appropriate to describe these reactions.  相似文献   

19.
An analysis is made of earlier work by the authors in investigating the catalytic activity of Fe(III) and Cu(II) ions in solution and on different surfaces in certain redox reactions: the decomposition of H2O2, oxidation of ascorbic acid, etc. General principles have been established which provide a means of predicting one or another action of sorption on the catalytic activity: if the reaction proceeds within the ion exchanger phase as in a homogeneous (concentrated) solution, it is most probable to have a decrease (or cessation) of the catalytic activities of the ions. If the ion catalysts are included in a surface complex in whose coordination sphere there are places which the substrate may occupy the reaction is facilitated. The analysis carried out of the influence of sorption on the catalytic activity of the iron and copper ions shows a favorable change in the catalytic activity for these ions during their sorption on different types of surfaces.Translated from Teoreticheskie i Éksperimental'naya Khimiya, Vol. 22, No. 6, pp. 706–711, November–December, 1986.  相似文献   

20.
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