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 IR MATIR spectra in the 1535-1680 cm-1 range were studied for epoxy-DPP resins (M N = 1650-3400) in coatings on germanium substrate obtained from oligomer solutions in methylene chloride and Cellosolve with the concentration c = 10-50%. The concentration dependences of the relative viscosity of narrow-MWD fractions of epoxy oligomers (M N = 1500-5300) in chloroform and Cellosolve solutions were studied. The structure of the network of cross-linked polymers based on epoxy (M N = 2100-3400) and phenol-formaldehyde (M N = 860) resins was studied by the electron-microscopic silver chloride decoration method. Based on the cluster lattice model, the optimal molecular weight and the concentration regimes were determined for epoxy oligomers in the lacquer composition for can protection.  相似文献   

2.
Kinetic viscosity curves (t) for solutions of a mixture of epoxy-4,4'-isopropylidenediphenol and phenol-formaldehyde resins in Cellosolve at 333 K were studied. The optimal precondensation time t was determined for the oligomer mixture in solution in the presence of orthophosphoric acid.  相似文献   

3.
Concentration dependence of isomer shift of K2SnCl6 and tin(IV) chloride was studied. Tin was hydrolyzed in both cation and anion exchange resins after water washing. Two pairs of these two solutions showing the same isomer shift values, and therefore the same , were selected for further studies in HCl solutions and in anion exchange resins. The first pair was 0.04 M K2SnCl6 and tin(IV) chloride solutions (δ=0.00 mm/s; =0): Their δ-[HCl] curves coincided each other. Tin was sorbed on anion exchange resin as pentachlorstannate(IV) at (HCl)≤6 N, and chiefly as hexachlorostannate(IV) at [HCl]≥7 N. The both complexes were sorbed when 6 N < [HCl] < 9 N. The δ-[HCl] curves of the second pair, 0.4 M K2SnCl6 and 3 M tin(IV) chloride(δ =0.23 mm/s; =3), also coincided each other, and tin was sorbed as pentachlorostannate(IV) from K2SnCl6 solutions in the HCl concentration range studied (≥5 N).  相似文献   

4.
The kinetics of the degradation of N-amino-3-azabicyclo[3,3,0]octane by chloramine has been studied by GC and HPLC in stoichiometric conditions in a solution buffered with NaOH/KH2PO4 and Na2B4O7.10 H2O between pH = 10.5 and 13.5. The second-order reaction exhibits specific acid catalysis which indicates competitive oxidation between the haloamine and the neutral and ionic forms of the bicyclic hydrazine. The enthalpy and entropy of activation were determined at pH = 12.89. In a nonbuffered solution, the interaction is autocatalyzed due to acidification of the mixture by the ammonium ions. In basic medium, the reaction forms an endocyclic hydrazone. A mathematical treatment based on an implicit equation allows a quantitative interpretation of all the phenomena observed over the above pH interval. This takes both the acid/base dissociation equilibria and the alkaline hydrolysis of the chloro-derivative into account. © 1995 John Wiley & Sons, Inc.  相似文献   

5.
The reaction of (tmeda)Pd(ClO4)2 (tmeda = N,N,N′,N′-tetramethylethylenediamine) with L (L = bis(4-(4-pyridylcarboxyl)phenyl)methane) affords the ionic cyclodimeric palladium(II) complex [(tmeda)Pd(L)]2(ClO4)4. The complex forms an unprecedented micro-sprout morphology via slow evaporation of acetone in a dilute concentration mixture of acetone and water without any template or additive. In contrast, the palladium(II) complex in a concentrated mixture forms uniform submicrospheres. The formation-process of the micro-sprout morphology has been explained in terms of a stepwise concentration effect. Furthermore, surface modifications and properties of the micro-sprouts via a typical anion exchange or sonication have been studied.  相似文献   

6.
The hydrothermal synthesis of analcime (ANA) with N,N′‐dibenzyl‐N,N,N′,N′‐tetramethylethylenediamine (DBTMED) as template was systematically studied. The various parameters that affect the crystallization of analcime, such as temperature, time, Al source, and Si/Al ratio were investigated. Systematic variations of these parameters revealed that ANA was obtained from the reaction mixture with the optimized ratios of SiO2/Al2O3 = 5–9.5 in presence of DBTMED, whereas template‐free clear solution methods require SiO2/Al2O3 ratio of greater than 25. When experiments were conducted at 130 and 150 °C for 4 days, a mixture of analcime and zeolite P was present in the samples, and a pure analcime sample could be obtained with heating in the temperature range 160–180 °C. When microwave and conventional heating were used, analcime could be obtained after 2 days. The obtained products were characterized by XRD, SEM, and IR spectroscopy.  相似文献   

7.
For the SN1, SN2, E1, E2, AND E1cb mechanisms exact solutions of the kinetic equations are compared with solutions obtained under the hypothesis of stationarity. Exact integrals are calculated numerically for a set of the essential reaction rate constants. Comparison with the stationary state solution demonstrates that both types of solutions are approximately equal in many cases if the transient particle concentration remains low, whereas the conventional requirement | \documentclass{article}\pagestyle{empty}\begin{document}$ \mathop x\limits^. $\end{document} | = 0 is unimportant. For approximation of exact kinetics by stationary kinetics, however, neither condition is sufficient nor necessary. Practical criterions for recognition of deviations from stationary kinetics are given.  相似文献   

8.
The dissociation, capacity, swelling, and water content of crosslinked methacrylic acid—methyl methacrylate resins have been measured. Resins were prepared with different degrees of crosslinking for the same carboxylate content, and vice versa. The ionic strength of the external solution was also varied, and the behavior of commercial resins compared with that of the laboratory resins. Potentiometric titration curves were obtained, and curves were also obtained by back-titration of the salt form of the resins with acid. The capacities showed that almost all carboxyl groups are accessible in resins containing 2.5% or 4.0% divinylbenzene, but not in those containing 8% or 12%. For these highly crosslinked resins the back-titration curves differed from the forward curves. Apparent dissociation constants pKa = pH + n log [(1 ? α)/α] decreased with increased ionic strength, increased with increased crosslinking, and showed no trend with carboxylate content. Swelling is decreased by increased salt concentration, particularly for lightly crosslinked resins. Maximum swelling is achieved at about 80% dissociation. The reciprocal of the swollen volume is proportional to the per cent of divinylbenzene. Commercial resins showed much lower swelling than laboratory prepared resins ostensibly having the same composition. The Gibbs-Donnan theory of resin dissociation was applied to calculate the intrinsic dissociation constant (pK0). Assuming a model of randomly kinked chains dissolved in the sorbed solution, good agreement with the expected value of 4.85 was found (calcd. pK0 = 4.81 = 0.14), except for the most highly crosslinked resins. For polyampholyte resins, agreement was found by using a model having a uniform potential distribution throughout the resin (pK0 = 4.9).  相似文献   

9.
The effect of concanavalin A on the structure of polymer hydrogels prepared via the free-radical copolymerization of acrylamide, N-(2-D-glucos)acrylamide, and N,N′-methylene-bis(acrylamide) is studied. When complexed with N-(2-D-glucos)acrylamide, concanavalin A is involved in copolymerization as a macromolecular crosslinking agent. This circumstance ensures a decrease of the degree of swelling of hydrogels in aqueous solutions with an increase in the concentration of concanavalin A in the initial monomer mixture. After the addition of glucose to an aqueous solution, the complex of concanavalin A with units of N-(2-D-glucos)acrylamide in the crosslinked copolymer dissociates and the degree of swelling of hydrogels increases considerably. Dissociation of the complex occurs at a strictly specified concentration of glucose in the solution that depend on the content of N-(2-D-glucos)acrylamide units in the copolymer. This phenomenon can be used for the controlled release of insulin previously introduced into the hydrogel through a change in the concentration of glucose in the solution.  相似文献   

10.
Cyclic voltammetry has been employed to investigate the mixed micellar behavior of the binary mixtures of different zwitterionic surfactants such as 3-(N,N-dimethylhexadecylammonio)propane sulfonate (HPS), 3-(N,N-dimethyltetradecylammonio)propane sulfonate (TPS) and 3-(N,N-dimethyldodecylammonio)propane sulfonate (DPS) with three triblock polymers (L64, F127 and P65) by using 2,2,6,6-tetramethyl-1-piperidinyloxy (TEMPO) as an electroactive probe at 25 °C. Critical micellar concentration (cmc) has been determined from the plots of variation in peak current (ip) versus the total concentration of surfactant/triblock polymer. Diffusion coefficient of the electroactive species has also been reported. The regular solution theory approximation has been used to determine various micellar parameters of ideal systems. The variation in micellar mole fraction (X1) of the zwitterionic surfactant supports the formation of mixed micelles, which are rich in triblock polymer component in the surfactant rich region of the mixture and vice versa. The regular solution interaction parameter (β) suggests the formation of mixed micelles due to the synergistic interactions in case of HPS/TPS/DPS + F127/P65 systems and gets affected by EO/PO ratio of triblock polymers.  相似文献   

11.
The kinetics of the hydrogen oxidation and the CO adsorption on a Pt (ultra)microelectrode is studied in a 0.5 M H2SO4 solution saturated with a mixture of gaseous H2 and CO at partial CO pressures p CO = 10–500 ppm. The balance between rates of diffusion and adsorption of CO at different adsorption times is studied. Studied is the effect of CO impurities in H2 on steady-state polarization curves for the hydrogen ionization and nonsteady-state curves of the oxidation current decay with time at 0.02–0.05 V. Conditions under which in a certain time interval and at a certain CO concentration the slope of an I vs. t curve is proportional to p CO are determined. The obtained dependence may be used when designing a technique for monitoring CO impurities in technical hydrogen.  相似文献   

12.
The kinetics and mechanism of the N2-N1-isomerization of 2-methoxycarbonyl-5-(p-X-phenoxy)- tetrazoles (X = H, CH3, NHCOCH3, Cl, Br, NO2) were studied by 1H NMR spectroscopy in a DMSO-d 6-CDCl3 mixture (25:75). The rate of isomerization of the N2-isomer into N1-isomer fit the first-order equation (after three half-conversion periods). The isomerization is accompanied by hydrolysis and decarboxylation. The Hammett plot of ln(k X k H) for the isomerization showed a good correlation with - values (- = 1.33, r = 0.965). A poor correlation with values was obtained. The kinetic data, the effect of solvent polarity, the substituent effects, and the results of AM1 quantum-chemical calculations suggest an ionic mechanism of the isomerization in polar solvents and a concerted mechanism in nonpolar solvents.  相似文献   

13.
In the micellar polymerization to prepare associative polyelectrolyte, the influence of solution environment on hydrophobical micro-block length has been investigated. The results revealed that the aggregation number (Nagg) of surfactant micelles and micro-block length (NH) of polymers closely related to electrolyte concentration in polymerization solution, and all of the surfactant and charged monomer concentration can change the free counterion concentration (Caq) to affect Nagg and NH. Therefore, the NH with traditional calculation method which Nagg is looked as constant value was not accurate enough. Subsequently, we studied the attribution of NH to rheological properties. The results indicated that the micro-block length can increase obviously the reversible network strength in solution to make a few contribution of thickening properties of associative polymer.  相似文献   

14.
A series of polylactide/poly(ethylene glycol) (PLA/PEG) block copolymers were synthesized by ring‐opening polymerization of L ‐ or D ‐lactide in the presence of mono‐ or di‐hydroxyl PEG. The effects of stereocomplexation on the physicochemical behavior of PLA/PEG copolymers in aqueous solution were investigated by varying the degree of stereocomplexation or PLLA/PEG to PDLA/PEG ratio. In mixture solutions of insoluble and soluble copolymers, stereocomplexation strongly affects the solubility of the copolymers. In mixture solutions of soluble copolymers, both the size and aggregation number (Nagg) of the aggregates vary as a function of the degree of stereocomplexation. It is suggested that the size variation of the aggregates with increasing the degree of stereocomplexation is dependent on Nagg changes which are determined by two effects: the self‐adjusting of the aggregates so as to minimize the free energy and thus to increase the Nagg, and the kinetics of aggregation which tend to form more aggregates and thus to decrease the Nagg. Combination of the two opposite effects well explains the diverse variations of Nagg and size of the aggregates as a function of the degree of stereocomplexation. © 2012 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys, 2012  相似文献   

15.
The electrochemical and chemical polymerization of acrylamide (AA) has been studied. The electrolysis of the monomer in N,N-dimethylformamide (DMF) containing (C4H9)4NClO4 as the supporting electrolyte leads to polymer formation in both anode and cathode compartments. The cathodic polymer dissolves in the reaction mixture and the anodic polymer precipitates during the course of polymerization. A plausible mechanism for the anodic and cathodic initiation reaction has been given. The chemical polymerization of acrylamide that has been initiated by HClO4 is analogous to its anodic polymerization. The polymer yield increases with an increase in concentration of the monomer and HClO4. Raising the reaction temperature also enhances the polymerization rate. The overall apparent activation energy of the polymerization was determined to be ca. 19 kcal/mole. The copolymerization of acrylamide was carried out with methyl methacrylate (MMA) in a solution of HClO4 in DMF. The reactivity ratios are r1 (AA) = 0.25 and r2 = 2.50. The polymerization with HClO4 appears to be by a free radical mechanism. When the polymerization of acrylamide is carried out with HClO4 in H2O, a crosslinked water-insoluble gel formation takes place.  相似文献   

16.
A series of chelating resins, derived from a macroreticular styrene-divinylbenzene (2%) copolymer beads grafted with various poly(ethylene glycols) HO? (? CH2? CH2? O? )n? H(n = 0, 4, 9, 13) and containing thiol groups as chelating functions, have been synthesized in a three-step reaction sequence. The structure of the functionalized resins was confirmed by IR spectrophotometry, elemental analysis, and differential scanning calorimetry. The complexation behavior of these thiol resins was investigated towards Hg(II), Cu(II), and Pb(II) ions in aqueous solution by a batch equilibration technique. The influence of pH on adsorption capacity was also examined. The adsorption values for metal ions' intake followed the order Hg(II) > Cu(II) > Pb(II). The affinity of these polymers towards Hg(II) ions was so high that the total mercury level in the liquid decreased from 20 ppm to below 10 ppb after 2 h of treatment. Polymers can be regenerated by washing with a solution of hydrochloric acid (6N) and 10% by weight of an aqueous solution of thiourea. © 1994 John Wiley & Sons, Inc.  相似文献   

17.
Reactions of NO in a Positive Streamer Corona Plasma   总被引:2,自引:0,他引:2  
The reaction of NO in a streamer corona plasma is studied systematically as a function of the composition of the gas mixture, the initial concentration of NO, and the discharge repetition rate. The experimental results show that the reactions of NO depend strongly on the composition of the gas mixture. Reduction is observed in the absence of oxidants such as oxygen and water, but at very high energy cost (>200 eV/NO). In the presence of both these oxidants, more than 90% of the NO conversion is oxidation. The lowest energy costs, 24 eV/NO for He mixtures and 45 eV/NO for N 2 mixtures, are obtained at water and oxygen concentration above 3% and at low discharge repetition rates (<10 Hz). Chemical kinetics calculations of the production of radicals in the plasma show a good agreement with the value derived from the experiments.  相似文献   

18.
Nucleophilic substitutions of Pd(N,N)Cl2[(N,N = 1-methyl-2-(arylazo)imidazole (RaaiMe), p-RC6H4N=NC3H2NN-1-Me; 2-(arylazo)pyridine (Raap), p-RC6H4N=NC5H4N; 2-(arylazo)pyrimidine (Raapm), p-RC6H4N=NC4H3N2 where R = H (a), Me (b), Cl (c)] with 8-quinolinol (HQ) have been examined by spectrophotometry at 298 K in MeCN solution. The product, Pd(Q)2, has also been confirmed by independent synthesis from Na2[PdCl4] and HQ in EtOH. The kinetics of the reaction have been studied under pseudo-first-order conditions and the analyses support a nucleophilic association path. A single phase reaction has been observed and follows the rate law, rate = a + k [Pd(N,N)Cl2] [HQ]2. Thus, the reaction is first order in [Pd(N,N)Cl2] and second order in [HQ]. External addition of Cl(LiCl) suppresses the rate. The rate increases as follows: Pd(RaaiMe)Cl2 < Pd(Raap)Cl2 < Pd(Raapm)Cl2.  相似文献   

19.
Summary The kinetics of the reaction between H2O2 and some Schiff base complexes of MnIII have been investigated in both aqueous and micellar sodium dodecyl sulphate (SDS) solution. The reaction rate is first order in both H2O2 and [complex], and inversely proportional to [H+]. The second-order rate constant increases in the sequence [Mn(salophen)(OAc)] > [Mn(salen)(OH2)]-ClO4 > [Mn(salen)(OAc)]H2O, where salen = N,N-bis-(salicylidene)ethylenediamine and salophen = N,N-bis-(salicylidene)-o-phenylenediamine. At SDS concentrations below the critical micellar concentration, there is almost no effect on the rate of reaction whereas at higher concentrations the reaction rate increases slightly. A mechanism involving MnII and a peroxo intermediate is proposed.  相似文献   

20.
The oxidative cyclization of the title compounds results in generally two different kinds of products. The first, 1-(N,N-bisacetylamino)-1,2,3-triazole 7 (R3 = CH3) is the primary product, while the second, 1-N-acetylamino-1,2,3-triazole 8 (R3 = CH3), when observed, is obtained via hydrolysis from the former during work-up and separation of the reaction mixture. The primary products are considered as resulting from intramolecular nucleophilic attack on the acetyl group, of the presumed zwitterionic intermediate 5 (R3 = CH3), by the N of the ambident N-acetylimine site of 5 .  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