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 265 毫秒
1.
Synthesis of poly(N-vinylformamide) (PNVF) and its subsequent hydrolysis to convert it to poly(vinyl amine) (PVAm) were performed. Kinetics of acidic and basic hydrolysis of poly(N-vinylformamide) (PNVF), and products of hydrolysis were investigated by using Fourier transform infrared, size exclusion chromatography, 1H NMR, and 13C NMR spectroscopies, and thermogravimetric analysis. It was observed that amide groups did not completely transform into amine groups by acidic hydrolysis of PNVF while the conversion of amides into amine groups via basic hydrolysis of PNVF was complete in 12 h, as confirmed by spectroscopic measurements. Results of extensive characterization revealed significant structural and conformational differences between acidic and basic hydrolysis products. Fluorescence spectroscopy was used for the first time to follow the conversion of amide groups into amine groups. The fluorescence intensity of PVAm obtained from basic hydrolysis of PVNF showed significant increase with amide/amine conversion. Finally, PVAm obtained from acidic hydrolysis of PNVF demonstrated potent antimicrobial activity, 10–20 times more, against common pathogens for example, C. albicans as fungal strain and E. coli, S. aureus, B. subtilis, and P. aeruginosa as bacterial strains as compared to PVAm obtained from basic hydrolysis.  相似文献   

2.
A convenient method of preparing chelating fiber with amine groups on the fiber surface was developed. The precursor polymer of Poly(N‐vinylformamide/acrylonitrile) (P(NVF/AN)) was synthesized via solution polymerization, using N‐vinylforaimde as a functional monomer. The solution of P(NVF/AN) was spun through a wet spinning method and the precursor fiber was hydrolyzed in the hydrochloric acid solution to convert formamide moieties to the corresponding amine. The influence of hydrolytic conditions on hydrolysis degree, such as hydrolysis temperature, hydrolysis time, and hydrochloric acid concentrations were examined experimentally. The hydrolysis degree of the precursor fiber was evaluated by potentiometric and conductometric titrations. The changes of the structure and properties of the fibers were characterized through infrared spectroscopy, scanning electron microscopy, and tensile strength tester. The results showed that the hydrolysis degree was limited in acidic hydrolysis because of the electrostatic repulsion among the cationic amine groups and proton. The hydrolysis degree of precursor fiber reached nearly 60%, and the chelating fiber remained the adequate mechanical properties under the suitable hydrolysis condition. © 2008 Wiley Periodicals, Inc. J Appl Polym Sci, 2008  相似文献   

3.
Poly(p‐phenylene benzobisoxazole) (PBO) fiber has received great interest because of its excellent mechanical properties and good thermal stability. The objective of this study was to expose degradation mechanism of PBO under neutral and acidic conditions by molecular mass and Fourier transform infrared (FTIR) spectroscopy. Results were not consistent with the classic degradation mechanism, which indicates that degradation should occur through the ring opening and chain scission of the benzoxazole ring. The FTIR absorption spectra of PBO suggested that the o‐hydroxy amide linkage (the open ring structure) was present in the PBO molecule chain to some extent because of the incomplete polymerization. Further investigation showed that hydrolysis might occur in the open ring section during hydrolytic degradation. Based on the experimental data, a new degradation mechanism was proposed. It suggests that, in the early and middle stages, hydrolysis occurred primarily in the o‐hydroxy amide linkage of the open ring. The concentration of the o‐hydroxy amide structure determined the speed of degradation of PBO. © 2011 Wiley Periodicals, Inc. J Appl Polym Sci, 2011  相似文献   

4.
Poly(vinylamine) (PVAm) was grafted on a poly(ethylene) (PE) film surface via the surface graft polymerization of N‐vinylformamide (NVF) and N‐vinylacetamide (NVA) and the subsequent hydrolysis of those grafted polymers. The surface was characterized by X‐ray photoelectron spectroscopy (XPS), contact angle, moisture absorption, and the leakage of electrostatic charge from the films. PNVF and PNVA were introduced onto the surface of the PE film successfully, in spite of the fact that the initiator for polymerization was a peroxide group. The grafted amounts of PNVF and PNVA were dependent on the grafting time. A PVAm‐grafted surface was obtained via the hydrolysis of the grafted PNVF. The grafted‐PNVA was not hydrolyzed under mild hydrolysis. The obtained PVAm‐grafted surface appeared to be useful for various applications, such as protein immobilization or chemical modification. © 1999 John Wiley & Sons, Inc. J Appl Polym Sci 72: 1583–1587, 1999  相似文献   

5.
p‐Perfluoro{1‐[2‐(2‐fluorosulfonyl‐ethoxy)propoxy]}ethylated poly(α‐ methyl styrene) 3 was synthesized via electron transfer of perfluoro‐di{2‐[2‐(2‐fluorosulfonyl‐ethoxy)propoxy]}propionyl peroxide 2 and poly(α‐methyl styrene) 1 at different reactant molar ratios (2 : 1). The modified polymer 3 was characterized by various techniques. The ring p‐substitution was proved by FTIR and 19FNMR. The desulfonation appeared at above 124°C was found by TGA. The molecular weight determined by GPC increased with the increase of reactant molar ratio, and the polydispersity values indicated there was no degradation of the parent polymer chain in the reaction. Followed by hydrolysis and acidification, the modified polymer 3 could be further quantitatively converted into its sulfonic form 4. Ion exchange capacity of novel polyelectrolyte 4 can be controlled by changing reactant molar ratio (2 : 1). © 2006 Wiley Periodicals, Inc. J Appl Polym Sci 100: 3615–3618, 2006  相似文献   

6.
王永丽  黄少斌 《应用化工》2005,34(8):517-518
合成聚天门冬氨酸阻垢剂时,都要经历中间产物聚琥珀酰亚胺(PSI)水解生成聚天冬氨酸钠盐这一步骤。研究探讨水解温度、反应时间、NaOH浓度和NaOH用量对聚琥珀酰亚胺水解产物聚天冬氨酸阻垢率的影响。通过正交试验,发现反应时间对水解产物聚天冬氨酸的阻垢率影响最大,水解温度影响较小;得到最佳水解条件为:水浴温度60℃,时间30 m in,NaOH浓度为2 mol/L,NaOH用量为20 mmol,即在此条件下制备的聚天冬氨酸阻垢率最高。根据实验数据,推测聚琥珀酰亚胺的水解动力学过程属于连串反应。  相似文献   

7.
Partially aminated poly‐N‐vinylformamides (APNVF) were prepared by the hydrolysis of PNVF and used as the retention aid of rosin size. The dual retention aids system, consisting of this modern polymer and aluminum sulfate (alum) for neutral‐alkaline paper sizing using acid rosin sizes, was evaluated by experiment. The results indicated that APNVF was very effective and a small amount of the polymer used together with alum considerably increased the size retention and sizing degree of paper under neutral‐alkaline conditions. The cationic charge density of APNVF significantly influenced the sizing efficiency of the rosin sizes. Furthermore, the retention of alkaline filler CaCO3 and paper strength were improved by the polymer addition. It is clear that the polymer can be used as a multifunctional additive for papermaking. © 2000 John Wiley & Sons, Inc. J Appl Polym Sci 78: 1805–1810, 2000  相似文献   

8.
Poly(vinyl acetate) (PVAc)-based particles are used in medical and biomedical applications, including vascular embolization (VE) procedures. However, to render the technique safer and reproducible, the hydrolysis of the microbeads must be performed in some applications in order to control the hydrophilicity of the material. For this reason, in the present work the alkaline hydrolysis of poly(vinyl acetate-co-methyl methacrylate)—P(VAc-co-MMA)—microbeads is investigated. The hydrolytic conversions of PVAc and poly(methacrylic acid) (PMMA) chains were evaluated in terms of the concentrations of methanol and acetate ions present in the reaction media, since these chemicals are by-products of alkaline hydrolyses of the ester groups of the polymers. It was observed that all investigated variables (temperature, time, and alkaline concentration) exert important effects on the final extents of hydrolysis, although the occurrence of nonlinear synergetic effects imposes the use of multivariable statistical analyses for proper characterization of variable effects. Besides, given the heterogeneous nature of the reaction system and the higher resistance to hydrolysis offered by PMMA segments, the overall hydrolytic conversions of the microbeads were low, as required in VE procedures. Finally, hydrolyses of P(VAc-co-MMA) chains can also lead to cleavage of polymer chains, as confirmed by reduction of average molar masses of the analyzed materials.  相似文献   

9.
A polyelectrolyte complex (PEC) was formed by mixing aqueous solutions of the polyanion poly(sodium phosphate) with the polycation poly(2‐vinylpyridinium chloride). Conductometric and potentiometric titrations indicated the electrochemical end point of each titration. In all cases the end point occurred at a unit molar ratio of polyanionic to polycationic groups that was approximately one. The stoichiometry was also confirmed by analysis of the supernatant liquid in conjunction with the weights of the initial components and complex. An analysis showed that the starting materials were regenerated after dissolution of the complex with a 2M HCl solution. The interaction of the bivalent cupric ions with PEC were also investigated. © 2002 Wiley Periodicals, Inc. J Appl Polym Sci 83: 3022–3028, 2002; DOI 10.1002/app.2332  相似文献   

10.
The synthesis of a thermally stable proton conducting polymer based on poly(phenylene oxide) (PPO) was carried out using 2,6‐dimethylphenol (DMP) and 2‐allylphenol (AP) as monomers. The copolymers using the two monomers were prepared with DMP to AP molar ratios of 20:80, 40:60, 60:40, and 80:20. The polymers and the copolymers were blended with poly(vinylidene fluoride) and cast as membranes. All the membranes were sulfonated and characterized for their suitability as a polymer electrolyte membrane. © 2002 Wiley Periodicals, Inc. J Appl Polym Sci 83: 1792–1798, 2002  相似文献   

11.
Poly(vinyl chloride) PVC pyrolysis and hydrolysis are conducted in a fixed bed reactor and in an autoclave, respectively, under different operating conditions such as the temperature and time. The product distribution is studied. For the PVC pyrolysis process, the main gas product is HCl (55% at 340°C), there is 9% hydrocarbon gas (C1–C5), the liquid product fraction is about 5% (at 340°C), and the solid residue fraction is about 31% (at 340°C). For the hydrolysis process, the main gas product is HCl (55.8% at 240°C) and the solid residue is about 49.6% (at 240°C). The pyrolysis liquid product is analyzed by using gas chromatography with magic‐angle spinning. Aromatic hydrocarbons are the main class (90%), of which the major part is benzene (33%). The residue produced through pyrolysis and hydrolysis is investigated by high‐resolution solid‐state 13C‐NMR. These details revealed by the high‐field NMR spectra provide importmant information about the chemical changes in the PVC pyrolysis and hydrolysis process. The mechanism of PVC hydrolysis dechlorination is also discussed. © 2003 Wiley Periodicals, Inc. J Appl Polym Sci 90: 3252–3259, 2003  相似文献   

12.
The crystallization kinetics of poly(butylene terephthalate) (PBT), poly(ethylene terephthalate) (PET), and their copolymers poly(1,4‐butylene‐co‐ethylene terephthalate) (PBET) containing 70/30, 65/35 and 60/40 molar ratios of 1,4‐butanediol/ethylene glycol were investigated using differential scanning calorimetry (DSC) at crystallization temperatures (Tc) which were 35–90 °C below equilibrium melting temperature . Although these copolymers contain both monomers in high proportion, DSC data revealed for copolymer crystallization behaviour. The reason for such copolymers being able to crystallize could be due to the similar chemical structures of 1,4‐butanediol and ethylene glycol. DSC results for isothermal crystallization revealed that random copolymers had a lower degree of crystallinity and lower crystallite growth rate than those of homopolymers. DSC heating scans, after completion of isothermal crystallization, showed triple melting endotherms for all these polyesters, similar to those of other polymers as reported in the literature. The crystallization isotherms followed the Avrami equation with an exponent n of 2–2.5 for PET and 2.5–3.0 for PBT and PBETs. Analyses of the Lauritzen–Hoffman equation for DSC isothermal crystallization data revealed that PBT and PET had higher growth rate constant Go, and nucleation constant Kg than those of PBET copolymers. © 2001 Society of Chemical Industry  相似文献   

13.
Poly(amide imide) copolymers were synthesized with different molar ratios of 4,4‐diphenylmethane diisocyanate, two types of aromatic dianhydrides (pyromellitic dianhydride (PMDA) and 3,3′,4,4′‐sulfonyl diphthalic anhydride (DSDA)), and a diacid, which was derived from 3,3′‐dinitrobenzidine and isophthaloyl chloride in a previous work. In this study, the copolymers were further reacted with a reducing agent, and the nitro groups in the copolymers were hydrogenated into amine groups. Then, the amine‐group‐containing poly(amide imide) copolymers were cyclized at 180°C to form the poly(benzimidazole imide amide) copolymers in poly(phosphoric acid), which acted as a cyclizing agent. The resultant copolymers were soluble in sulfuric acid and poly(phosphoric acid) at room temperature and in sulfolane or N‐methyl‐2‐pyrrolidone under heating to 100°C with 5% lithium chloride. According to wide‐angle X‐ray diffraction, all the copolymers were amorphous. According to thermal analysis, the glass‐transition temperature ranged from 270 to 322°C, and the 10% weight‐loss temperature ranged from 460 to 541°C in nitrogen and from 441 to 529°C in air. © 2003 Wiley Periodicals, Inc. J Appl Polym Sci 91: 378–386, 2004  相似文献   

14.
A new type of unsaturated poly(ester‐amide) viz maleic anhydride‐phthalic anhydride‐ethylene glycol‐neopentylene glycol‐glycin copolymer was prepared by melt polycondensation. The copolymer was characterized by FT‐IR, gel permeation chromatography, and thermal gravimetric analysis. The molecular structure of crosslinked unsaturated poly(ester‐amide) was determined by wide‐angle X‐ray diffraction. Hydroxyapatite (n‐HA) was used to boost up the new unsaturated poly (ester‐amide), the flexural properties of n‐HA/unsaturated poly(ester‐amide) nanocomposites with different n‐HA content were measured. Studies of degradation behavior were carried out in simulated body fluid at pH 7.4 and 37°C, the flexural strength changes and cumulative mass loss of n‐HA/ unsaturated poly(ester‐amide) nanocomposites were measured at different degradation times. The n‐HA/unsaturated poly(ester‐amide) nanocomposites was hydrolyzed in 1M NaOH standard solution at room temperature to study the mass loss with different n‐HA contents. All the preliminary results suggested that n‐HA/unsaturated poly(ester‐amide) nanocomposites might be potentially used as a new type of bone fixation material. © 2011 Wiley Periodicals, Inc. J Appl Polym Sci, 2011  相似文献   

15.
This work investigated membrane formation using Lewis acid–base complexes in a polymer solution, which consisted of poly(ether sulfone) (PES), Lewis acid–base complexes formed by N‐methyl‐2‐pyrrolidone (NMP, Lewis base), and dicarboxylic or monocarboxylic acids from a homologous series (Lewis acids). The solutions were characterized by viscosity measurements, IR spectroscopy, cloud point determination, and light transmission experiments. The membranes were characterized by scanning electron microscopy and gas permeation tests. The results indicated that the solvent–additive interaction, which is a function of their capacity to form complexes, and the acid chain length directly affect the viscosity and miscibility region. Consequently, these parameters combined with the complex dissociation influence the precipitation velocity of the polymer solutions, which will then affect the membrane transport properties. It is also pointed out that the membranes prepared by using 25 wt % PES at the same acid/NMP molar ratios and with different acids presented permeability coefficients in agreement with the binodal shift obtained in pseudoternary phase diagrams. Furthermore, when these solutions were exposed to the environment for a long period of time, the demixing onset sequence also agreed with the miscibility region for all solutions, except for the adipic acid solution because of its extremely high viscosity. © 2002 Wiley Periodicals, Inc. J Appl Polym Sci 83: 2022–2034, 2002  相似文献   

16.
Cationic polymer charge and polymer degradability each play a crucial role for packaging and delivering plasmid DNA. High density cationic charge has been shown to enhance transfection efficiency but may give rise to undesirable toxicity. Polyvinylamine (PVAm) nanogels bearing discrete amounts of surface charge were used to systematically examine the balance between transfection efficiency and cytotoxicity. Poly(N‐vinylformamide) (PNVF) nanogels were prepared via an inverse emulsion polymerization reaction and crosslinked with a nondegradable or acid‐labile crosslinker. The nanogels were then hydrolyzed to yield varying degrees of primary amines. The degree of conversion from PNVF to PVAm was controlled using different concentrations of NaOH and hydrolysis times. PVAm nanogel size and charge ranged from 150 to 310 nm, and +3.5 to +18 mV, respectively. These cationic particles were then complexed with pDNA encoding for luciferase. The cytotoxicity of PVAm nanogels and the transfection efficiency of PVAm/DNA complexes were evaluated in carcinomic human alveolar basal epithelial cells (A549). The cytotoxicity of PVAm nanogels increased with increasing accessible charge as expected. Transfection efficiency increased with increasing amounts of amine groups for nondegradable nanogels. Interestingly, acid‐labile nanogels bearing low charge demonstrated more sustained gene transfection when compared with the more highly charged nanogels. These observations suggested that the use of degradable particles with less charge may reduce cytotoxicity without compromising overall transfection efficiency. © 2010 Wiley Periodicals, Inc. J Appl Polym Sci, 2010  相似文献   

17.
The imidization of poly(styrene‐co‐maleic anhydride) with amines may improve some of its end‐use properties. The objective of this study was to examine the mechanism and kinetics with aniline (ANL) as an amine of the preparation of poly(styrene‐coN‐phenyl maleimide). The reaction was carried out in a tetrahydrofuran solution at 25–55°C and in an ethylbenzene solution at 85–120°C. The extent of the reaction was determined by conductance titration, a new and simple method. Two consecutive reactions were involved in the imidization: ring opening to produce an acido‐amide group and ring closing to form a corresponding imide group. The imidization rate was greatly influenced by the reaction temperature and the molar ratio of ANL to the anhydride. A model for the imidization kinetics over a wide range of reaction temperatures and concentration ranges was developed and validated. © 2006 Wiley Periodicals, Inc. J Appl Polym Sci 100: 2744–2749, 2006  相似文献   

18.
2,6‐Diphenoxybenzonitrile (DPOBN) was synthesized by reaction of phenol with 2,6‐difluorobenzonitrile in N‐methyl‐2‐pyrrolidone in the presence of KOH and K2CO3. Poly(aryl ether ketone ketone)/poly(aryl ether ether ketone ketone) copolymers with pendant cyano groups were prepared by the Friedel–Crafts electrophilic substitution reaction of terephthaloyl chloride with varying mole proportions of diphenyl ether and DPOBN using 1,2‐dichloroethane as solvent and N‐methyl‐2‐pyrrolidone as Lewis base in the presence of anhydrous AlCl3. The resulting polymers were characterized by various analytical techniques, such as FT‐IR, differential scanning calorimeter, thermal gravimetric analysis, and wide‐angle X‐ray diffraction. The crystallinity and melting temperature of the polymers were found to decrease with increase in concentration of the DPOBN units in the polymer. Thermogravimetric studies showed that all the polymers were stable up to 514°C in N2 atmosphere. The glass transition temperature was found to increase with increase in concentration of the DPOBN units in the polymer when the molar ratios of DPOBN to DPE ranged from 10/90 to 30/70. The copolymers containing 30–40 mol % of the DPOBN units exhibit excellent thermostability at (350 ± 10)°C and have good resistance to acidity, alkali, and organic solvents. © 2007 Wiley Periodicals, Inc. J Appl Polym Sci 104: 3601–3606, 2007  相似文献   

19.
The phenomenon of enthalpic relaxation was evaluated for poly(lactide‐co‐glycolide) (PLGA, 50:50), in terms of storage of nanospheres for use as a controlled drug delivery system. Samples were stored for different times and temperatures below the glass transition temperature (Tg). Relaxation occurred at a significant rate up to 15 degrees below the Tg of 39.2°C. The effect of polymer morphology was considered by comparing the relaxation kinetics of the raw polymer with that of nanospheres formed using a novel technique. The nanospheres were shown to have a larger change in heat capacity at the glass transition and a longer average relaxation time than that of the raw polymer, and the relationship between these two parameters was discussed. For both the raw polymer and the nanospheres, relaxation was found to occur at a significant rate at room temperature. The storage of this system at subambient temperatures was therefore deemed important for maintaining the physicochemical properties of the system. © 2002 Wiley Periodicals, Inc. J Appl Polym Sci 86: 1868–1872, 2002  相似文献   

20.
Samples of polyglycolide were degraded in various buffered solutions in order to investigate the effects of buffer concentration, pH and particular buffer ions on the hydrolysis reaction. The small‐angle X‐ray scattering profiles changed with degradation; long periods were calculated and their changes with degradation used as a measure of the degradation rate. The release of a model drug, theophylline, was also investigated under these different buffering conditions. In more concentrated buffer solutions theophylline was released slightly earlier and the samples began to swell sooner. Increasing the pH of the phosphate buffer also resulted in an earlier swelling and release of drug. These effects were attributed to the neutralisation and increased solubility of acidic degradation products in more concentrated or alkaline buffer solutions. When an imidazole–HCl rather than a phosphate or citrate–citric acid buffer system was used, swelling and drug release were accelerated, as imidazole catalysed the ester hydrolysis reaction. The results are interpreted within the context of a four‐stage model of PGA degradation. © 2003 Society of Chemical Industry  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