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.
Anion exchange membranes containing N‐isopropylacrylamide as a component were prepared, and their electrochemical properties were examined. The membranes were crosslinked with ethylene glycol dimethacrylate and contained weakly basic or strongly basic anion exchange groups. The dependence of electrochemical properties of the membranes (electrical resistance, transport number of anions, water content, and reduced osmotic flux) on temperature was completely different from those of the anion exchange membrane without N‐isopropylacrylamide. For example, the reduced osmotic flux decreased with increasing temperature until 40°C, and the transport number of chloride ions increased with increasing temperature from 25.0°C, although those of the conventional membrane monotonously increased or decreased. The transport numbers of various anions relative to chloride ions in electrodialysis were evaluated at a different temperature. Although the transport numbers between anions did not change appreciably in the conventional membrane with temperature, those of the anion exchange membranes with N‐isopropylacrylamide changed with a temperature dependent on the hydration degree of anions: permeation of less‐hydrated anions such as nitrate and bromide ions compared with chloride ions increased with increasing temperature, and that of strongly hydrated anions such as sulfate and fluoride ions decreased with increasing temperature. This is based on the increase or decrease in uptake of the anions in the membrane with the change in temperature because hydrophilicity of the membranes changes with temperature due to the apparent aggregation of isopropyl groups in the membranes. And the change in electrochemical properties and transport numbers of various anions relative to chloride ions with temperature was completely reversible with increasing or decreasing temperature. © 1999 John Wiley & Sons, Inc. J Polym Sci B: Polym Phys 37: 793–804, 1999  相似文献   

2.
The new monomers 3-(3-methacryloxypropoxy)benzylamine and 4-(3-methacryloxypropoxy)benzylamine in the form of hydrochlorides have been synthesized and radically copolymerized under various conditions with comonomers of different hydrophilicity, including N-acryloylmorpholine, N-acryloylpyrrolidine, N,N-dimethylacrylamide, and N-vinylbenzoylmorpholine, to obtain linear soluble, granular crosslinked, and silica-based macromolecular systems designed for investigating the action mechanism of benzylamine oxidase. Some characteristics of the prepared materials, including scanning electron microscopy photographs, are reported. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 3109–3118, 1999  相似文献   

3.
Several N‐(substituted phenyl) citraconimides containing phenolic hydroxyl groups (I) were prepared. I were esterified with acryloyl chloride producing the corresponding acrylate esters (II). II were free radically polymerized yielding linear polyacrylates (III). The citraconimidyl vinyls did not participate in the polymerization. The resulting polymers (III) were cured thermally or through the crosslinking agent N,N‐(p‐phenylene)dimaleimide. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 427–433, 1999  相似文献   

4.
The anionic polymerization of N‐acryloyl‐N′‐(4‐methylbenzoyl)urea (1) was carried out at 80°C for 24 h in DMF, DMSO, acetonitrile, or toluene by t‐BuOK or DBU (3 mol %) as an initiator to obtain polymer 3 in a good yield. The structure of 3 was dependent upon the initiator used, in which t‐BuOK selectively conducted the hydrogen‐transfer polymerization, while DBU partially induced the vinyl polymerization (16–20%). Likewise, N‐acryloyl‐N‐methyl‐N′‐(4‐methylbenzoyl)urea (2, i.e., an N‐methylated derivative of 1) was subjected to the hydrogen‐transfer polymerization. Although the yield of the polymer was lower in comparison with 1, the structure of the obtained polymer 4 was similarly governed by the initiator. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 465–472, 1999  相似文献   

5.
Terpolymers composed of Nn‐propylacrylamide (NPAAm), butyl methacrylate (BMA), and N,N‐diethylaminoethyl methacrylate (DEAEMA) were prepared in an attempt to investigate the temperature‐induced phase transition and its mechanism. Poly(NPAAm) showed the lower critical solution temperature (LCST) around 24°C in water. With the incorporation of DEAEMA with NPAAm, the LCST change was characterized by an initial increase. However, the LCST was shifted to the lower temperature at the later stage. This might be explained in terms of hydrophilic/hydrophobic contribution of DEAEMA to the LCST. The swelling behavior of copolymer gel in the various solvents and spin‐lattice relaxation time (T1) study by NMR strongly suggested the hydrophilic/hydrophobic contribution of DEAEMA to the LCST depending on the local environment. © 1999 John Wiley & Sons, Inc. J Polym Sci B: Polym Phys 37: 1407–1411, 1999  相似文献   

6.
The synthesis and characterization of a series of novel poly(aryl amide imide)s based on diphenyltrimellitic anhydride are described. The poly(aryl amide imide)s, having inherent viscosities of 0.39–1.43 dL/g in N-methyl-2-pyrrolidinone at 30°C, were prepared by polymerization with aromatic diamines in N,N-dimethylacetamide and subsequent chemical imidization. All the polymers were amorphous, readily soluble in aprotic polar solvents such as DMAC, NMP, dimethylsulfoxide, N,N-dimethylformamide, and m-cresol, and could be cast to form flexible and tough films. The glass transition temperatures were in the range of 284–366°C, and the temperatures for 5% weight loss in nitrogen were above 468°C. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 4541–4545, 1999  相似文献   

7.
N-Methylaniline-, diphenylamine-, and N-phenylnaphthylamine-blocked toluene diisocyanates (TDI) were prepared and characterized by IR, NMR spectroscopy, and nitrogen content analyses. The structure–property relationship of these adducts was established by reacting with hydroxyl-terminated polybutadiene (HTPB). The cure rate of the adduct increases from the N-phenylnaphthylamine- to diphenylamine- and to N-methylaniline-blocked TDI adduct. Simultaneous TGA/DTA results also confirm this trend, and the thermal stability of the adduct decreases in the following order: N-phenylnaphthylamine–TDI > diphenylamine–TDI > N-methylaniline–TDI. The gas chromatogram of the amine-blocked isocyanate confirms that the thermolysis products are the blocking agent and isocyanate. The solubilities of the adducts were carried out in polyether, polyester, and hydrocarbon polyols, and it was found that the N-methylaniline–TDI adduct shows higher solubility than the rest and also found that the polyester polyol shows higher solvating power against the adducts than the polyether and hydrocarbon polyols. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 1815–1821, 1999  相似文献   

8.
The fluorescence behavior of 1,1′‐dimethyl‐2,2′‐carbocyanine and pN,N‐dimethylamino‐styryl‐2‐ethylpyridinium was investigated in several epoxy systems. Time‐correlated single photon counting was used for all fluorescence measurements to obtain the rate constant for viscosity or mobility‐dependent nonradiative processes of the probe. Microviscosity effects were discussed on the basis of a model describing the microfriction between matrix and probe molecules. The probes investigated are able to detect the glass‐transition temperature of the materials investigated. These compounds also show a dependence on the mobility in the glassy state. The probes applied in this work also can be used to monitor the crosslinking process of several epoxy systems containing 4,4′‐diaminodiphenylmethane (DDM) as curing agent. The epoxides used for the crosslinking process were 2,2′‐[(1‐methylethylidene)bis(4,1‐phenyleneoxymethylene)bis‐oxiranemethaneamine] [common name, diglycidyl ether of bisphenol A (DGEBA)], N‐oxiranylmethyl‐N‐phenyl‐oxiranylmethane [common name, diglycidyl aniline (DGA)], and epoxy novolacs of different functionality. The networks obtained have a different morphology, which was studied by the fluorescence probe technology. The structure of the epoxy compound has an important influence on the probe behavior because both network density and size of the free volume influence the photochemical behavior of the probe. © 1999 John Wiley & Sons, Inc. J Polym Sci B: Polym Phys 37: 1367–1386, 1999  相似文献   

9.
An ABA triblock copolymer of polyvinyl acetate-b-polystyrene-b-polyvinyl acetate (PVAc-PS-PVAc) was successfully synthesized with a binary system composed of polystyrene with N,N-dimethylaniline end groups (PSda) and benzophenone to initiate the polymerization of vinyl acetate under UV irradiation. The PSda was obtained by capping the living polystyrene macrodianion with p-(dimethylamino) benzaldehyde in excess. The PVA-PS-PVA could then be obtained by hydrolysis of PVAc-PS-PVAc in the sodium ethoxide benzene solution. The intermediates and desirable copolymers were characterized by GPC, IR, and 1H-NMR in detail. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 2595–2600, 1999  相似文献   

10.
N‐Stearylacrylamide (SAAm), N‐oleylacrylamide (OAAm), and N‐laurylacrylamide (LAAm) were synthesized. They were characterized by 1H‐NMR, 13C‐NMR, FT‐IR, melting point measurements, and elemental analysis. The copolymerizations of SAAm, OAAm, and LAAm with 2‐[(3‐(acrylamido)propyl)dimethylammonio]ethyl 2′‐isopropyl phosphate were carried out, and a series of amphiphilic poly(acrylamide)s (1a,b, 2, and 3a,b) were obtained. These copolymers showed polyelectrolyte behavior in their viscous properties in polar solvents. X‐ray diffraction analysis indicated that the copolymers 1a,b formed similar stacked bilayers with hydrophilic groups and hydrophobic parts. The polymorphic phase transition of these copolymers was also observed by DSC. In addition, the monolayers as well as LB films of these amphiphilic copolymers were prepared on the surface of water and their π–A isotherms were investigated at different temperatures. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 1293–1302, 1999  相似文献   

11.
Copolymers of 4-vinylpyridine (4VP) or N,N′-diethylaminoethyl methacrylate (DEAEMA) were synthesized with various comonomer feed ratios (30 : 70, 50 : 50, and 70 : 30). Charge densities along the polymeric chain were changed in a pH-dependent manner. The copolymers induced membrane fusion of negatively charged liposomes followed by aggregation in a pH-dependent manner, which was investigated by resonance energy transfer. It was found that pH-dependent fusion of liposomes was dependent on the composition of 4VP and DEAEMA in the copolymer with the degree of protonation proportional to the difference in pKa value of the two monomers. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 2305–2309, 1999  相似文献   

12.
Rate constants and activation parameters for decrystallization of Avicel PH-101 cellulose, and bagasse-based cellulose in presence of LiCl/N,N-dimethylacetamide solvent system have been determined from dependence of the index of crystallinity of cellulose, Ic, on time, under nonisothermal conditions. Calculated rate constants and activation parameters are negligibly dependent on the degree of polymerization of the natural polymer. Under experimental conditions used, derivatization of cellulose can be started after 3 h of cellulose–solvent contact. The relevance of our results to the industrial application of derivatization under homogeneous solution conditions is discussed. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 3738–3744, 1999  相似文献   

13.
In this communication, we first used [60]fullerene as initiator to initiate the bulk polymerization of N-vinylcarbazole (NVC) monomer at 70°C (slightly higher than the melting point temperature, 65°C, of NVC). A reasonable polymerization reaction pathway via C60-NVC ion-radical pairs is suggested. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 3745–3747, 1999  相似文献   

14.
The reaction-induced phase separation in a tetrafunctional epoxy–cyclic anhydride system modified with polysulfone (PSF) was followed by optical microscopy (OM), light scattering (LS), and scanning electron microscopy (SEM). The selected system was N,N,N′,N′-tetraglycidyl-4,4′-diamino diphenylmethane cured with methyl tetrahydrophthalic anhydride, in the presence of variable PSF concentrations. The different experimental techniques allow us to establish the phase separation mechanism. For modifier concentrations close to the critical point, 10 and 15 wt % PSF, phase separation proceeded by spinodal demixing (SD). For a modifier concentration much lower than the critical point, 5 wt % PSF, phase separation occurred via the nucleation and growth (NG) mode. For 7.5 wt % PSF, depending on the cure temperature, SD or NG was observed. Dynamic mechanical behavior of the resulting materials had been discussed based on fractionation of different species during the phase separation process. The fracture toughness increased significantly when bicontinuous (10 wt % PSF) or phase-inverted (15 wt % PSF) structures were generated. For mixtures containing 15 wt % PSF, the dependence of fracture toughness on the stoichiometric ratio (anhydride groups/epoxy groups) was analyzed. © 1999 John Wiley & Sons, Inc. J Polym Sci B: Polym Phys 37: 2711–2725, 1999  相似文献   

15.
([bis[μ-[(2,3-butanedione dioximato)(2-)-O:O′]] tetrafluorodiborato(2-)-N,N′,N″,N‴] cobalt), CoBF, has been used for the effective catalytic chain transfer of alkyl methacrylate homo- and copolymers under emulsion polymerization conditions. The catalytic chain transfer process reduces the rate of polymerization such that when the monomer is fed over 60 min the instantaneous conversion is low enough for the particle to be swollen with monomer, allowing diffusion of the catalysts between the aqueous and monomer phases. When the amount of the catalyst is reduced, the rate is increased, eventually leading to viscous, glassy particles that prevent catalyst mobility, which is observed as a breakdown in the polymerization mechanism. This can be circumvented by the addition of a 20% shot of monomer at the start of the reaction. The effective chain transfer coefficient decreases on increasing the length of the ester group of the methacrylate. The analysis of the polymers made by the technique described shows that the Tg of the polymers observe a broad transition due to the effect of chain length being pronounced at low molecular mass. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 3549–3557, 1999  相似文献   

16.
Optically active 2,2′-bis(2-trifluoro-4-aminophenoxy)-1,1′-binaphthyl and its corresponding racemate were prepared by a nucleophilic substitution reaction of 1,1′-bi-2-naphthol with 2-chloro-5-nitrotrifluorotoluene and subsequently by the reduction of the resulting dinitro compounds. A series of optically active and optically inactive aromatic polyimides also were prepared therefrom. These polymers readily were soluble in common organic solvents such as pyridine, N,N′-dimethylacetamide, and m-cresol and had glass-transition temperatures of 256 ∼ 278 °C. The specific rotations of the chiral polymers ranged from 167 ∼ 258°, and their chiroptical properties also were studied. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 4536–4540, 1999  相似文献   

17.
A novel polyaryloxydiphenylsilane was synthesized successfully by solution polycondensation of 2,2′-dimethyl-biphenyl-4,4′-diol with diphenyldichlorosilane and the catalyst triethylamine in toluene at 80 °C. Polymers with a relatively high inherent viscosity and yield were obtained when the reactions were carried out in aromatic and lipophilic solvents. The novel polyaryloxydiphenylsilane was soluble in chlorinated aliphatic hydrocarbons such as methylene chloride and chloroform as well as in polar solvents such as dimethyl sulfoxide, N,N-dimethylformamide, and N,N-dimethylacetamide and also in some common organic solvents such as benzene and toluene. However, it was insoluble in both aliphatic hydrocarbons as well as in alcoholic solvents. The polyaryloxydiphenylsilane began losing weight around 400 °C under a nitrogen atmosphere, and the 10% weight-loss temperature was 473 °C. The glass-transition temperature of the polyaryloxydiphenylsilane was 102 °C. The glass transition could be lowered by the copolymerization technique with 2,2-bis(4-hydroxy-3,5-dimethylphenyl)propane as an aromatic diol comonomer. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 4591–4595, 1999  相似文献   

18.
A series of novel soluble poly(amide-imide)s were prepared from the diimide-dicarboxylic acid, 2,2-bis[N-(4-carboxyphenyl)-phthalimidyl]hexafluoropropane, with various diamines by the direct polycondensation in N-methyl-2-pyrrolidinone containing CaCl2 using triphenyl phosphite and pyridine as condensing agents. All the polymers were obtained in quantitative yields with inherent viscosities of 0.78–1.63 dL g−1. The polymers were amorphous and readily soluble in aprotic polar solvents such as N-methyl-2-pyrrolidinone, N,N-dimethylacetamide, N,N-dimethylformamide, and dimethyl sulfoxide as well as in less polar solvents such as pyridine and γ-butyrolactone, and also in tetrahydrofuran. The polymer films had tensile strength of 84–129 MPa, an elongation at break range of 6–22%, and a tensile modulus range of 2.0–2.7 GPa. The glass transition temperatures of the polymers were determined by DSC method and they were in the range of 240–282°C. These polymers were fairly stable up to a temperature around or above 400°C, and lose 10% weight in the range of 450–514°C and 440–506°C in nitrogen and air, respectively. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 2629–2635, 1999  相似文献   

19.
Regiocontrolled polymer (2) having 2-naphthol unit was prepared by oxidative coupling polymerization of bis(2-naphthol) (1). Polymerizations were conducted in dichloromethane in the presence of [di-μ-hydroxo-bis(N,N,N′,N′-tetramethylethylenediamine)copper(II)] chloride [CuCl(OH)TMEDA] under air at room temperature, producing polymers with number-average molecular weights up to 12,000. The structure of polymer 2 was characterized by 270 MHz 1H–NMR and 68.5 MHz 13C–NMR spectroscopies and was estimated to consist almost completely of 1,1′-linkage. The polymer was readily soluble in polar aprotic solvents and tetrahydrofuran at room temperature. Thermogravimetric analysis of polymer 2 showed 10% weight loss at 450°C in nitrogen. The model reactions were studied to clarify the applicability of CuCl(OH)TMEDA for coupling of naphthol derivatives. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 3702–3709, 1999  相似文献   

20.
A series of new poly(ether imide)s containing the naphthalimide moiety were prepared from bis(4-fluorobenzoyl)naphthalimides and several bisphenols by aromatic nucleophilic displacement polymerization. These polyimides had inherent viscosities in the range of 0.31–1.04 dL/g in chloroform and glass transition temperatures of 283.0–341.6°C by differential scanning calorimetry. The onset temperature for 5% weight loss for all the polymers was over 448°C, as assessed by thermogravimetry at a heating rate 10°C/min in nitrogen. In addition, these novel polyimides exhibited good solubility in organic solvents including N-methyl-2-pyrrolidone, N,N-dimethylformamide, N,N-dimethylacetamide, 1,1,2,2-tetrachloroethane and chloroform. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 3227–3231, 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