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 purpose of this study was to investigate the effects of reaction pH conditions on thermal behavior of urea–formaldehyde (UF) resins, for the possible reduction of formaldehyde emission of particleboard bonded with them. Thermal curing properties of UF resins, synthesized at three different reaction pH conditions, such as alkaline (pH 7.5), weak acid (pH 4.5), and strong acid (pH 1.0), were characterized with multiheating rate method of differential scanning calorimetry. As heating rate increased, the onset and peak temperatures increased for all three UF resins. By contrast, the heat of reaction (ΔH) was not much changed with increasing heating rates. The activation energy (Ea) increased as the reaction pH decreased from alkaline to strong acid condition. The formaldehyde emission of particleboard was the lowest for the UF resins prepared under strong acid, whereas it showed the poorest bond strength. These results indicated that thermal curing behavior was related to chemical species, affecting the formaldehyde emission, while the poor bond strength was believed to be related to the molecular mobility of the resin used. © 2006 Wiley Periodicals, Inc. J Appl Polym Sci 100: 422–427, 2006  相似文献   

2.
Observations and conclusions drawn from a model N-methylol reactant system have been used to study formaldehyde release from cotton fabric treated with N,N′-dimethylol-4,5-dihydroxy-ethyleneurea (DMDHEU). Reagent residues produced by DMDHEU in the presence of sodium bisulfate or zinc nitrate were found to be more complex than those formed from the N-methylolpyrrolidone (NMP) reactant system, and the exact nature of some of the residues from DMDHEU could not be established. Zinc nitrate produces higher fixation of DMDHEU on cotton as compared to sodium bisulfate and also reduces formaldehyde release. As is the case with the NMP system, the lower amount of formaldehyde release with zinc nitrate catalyst is believed to be at least partially due to the formation of a complex between the N-methylol reactant, DMDHEU, and zinc nitrate. The C? OCell bonds in both finishes resist acid-catalyzed cleavage but cleavage of the N? C bonds in N-methylol groups (to release formadehyde) in the presence of acid is greater for the DMDHEU system than for the NMP system. Results with the DMDHEU and NMP systems are similar, but because of the complex nature of the products formed in the DMDHEU system it is very difficult to correlated formaldehyde release with specific chemical structures of residues or reactants.  相似文献   

3.
Differential scanning calorimetry was used on a range of synthesized phenol–formaldehyde (PF) resols to discover relationships between formulation parameters or physical properties of resols, and their thermal behavior during cure. The thermograms showed either one or two exothermic reactions. The lower exothermic peak temperature varied between 98 and 129°C with changes in the free formaldehyde content. This exotherm is caused by the addition of free formaldehyde to phenolic rings. The upper exothermic peak temperature varied from 139 to 151°C, with the higher temperatures occurring when the formaldehyde-to-phenol molar ratio was low or the total amount of sodium hydroxide relative to phenol was high. These two factors led to resins which contain a somewhat higher level of unreacted ortho or para aromatic ring positions and no free formaldehyde. Consequently, condensation is probably not solely by the faster self-condensation through hydroxymethyl groups, but also includes the slower condensation of hydroxymethyl groups with unreacted ring positions. Gel times show trends with changes of formulation parameters somewhat similar to trends of the upper exothermic peak temperatures.  相似文献   

4.
A series of poly-secondary aromatic amines was prepared by the acid-catalyzed condensation of N-methylaniline with formaldehyde. These condensation products were evaluated as curing agents for a standard epoxy prepolymer, namely bis(2,3-epoxypropoxyphenyl)propane. Various molar ratios of N-methylaniline to formaldehyde were used to produce polyamine condensation products with varying degrees of polymerization. These molar ratios ranged from 1.60 to 1.30. Dielectric properties of the cured epoxy resins at temperatures above 100°C. were highly dependent upon these ratios. Lower-ratio condensation products produced resins with optimum dielectric properties at the higher temperature and also with good physical properties.  相似文献   

5.
Formation and composition of trans fatty acids (TFA) in corn oil during the frying process at different temperatures was investigated using an improved analytical gas chromatography (GC) method. At temperatures above 180 °C and heating times over 2 h there was a significantly increased TFA content, which increased further at higher temperatures. The amounts of trans C18:1 and trans C18:2 increased from 0.64 and 9.0 to 7.86 and 65.48 mg/g after heating at 260 °C for 12 h, while the total amount of TFA increased from 9.64 to 73.34 mg/g. However, the contents of linoleic acid and α-linolenic acid under the same conditions reduced from 460.3 and 7.4 to 209.8 and 1.8 mg/g, respectively. These results show that heating can significantly induce cis to trans isomerization of unsaturated triglycerides, particularly at high temperatures.  相似文献   

6.
The effects of time and temperature on dynamic viscoelastic properties of Chinese fir (Cunninghamia lanceolata [Lamb.] Hook) were investigated using dynamic mechanical analysis in this study. The isothermal tests were applied to the small clear specimens with a moisture content of about 0.6% at constant temperatures ranging from 25 to 200°C for 550 min at atmospheric pressure. Changes in storage modulus and loss tangent with heating time were examined. The results indicated that heating time mainly resulted in thermal softening, thermal degradation of wood, and the reduction of wood stiffness. At more than 60°C, the reduction in storage modulus was accelerated generally as the wood was subjected to a higher temperature or longer heating time. At constant temperatures of 140 and 160°C, a relaxation phenomenon was observed with a slight change in weight, which could be attributed to the relocation of lignin molecules. At the temperature range of 140 to 180°C, the higher the heating temperatures, the earlier the tanδ peak appeared. It is suggested that the wood thermal softening occurs at higher temperatures with shorter heating times or at lower temperatures with longer heating times. At temperatures of 180 and 200°C, the loss of amorphous polysaccharides due to thermal degradation is considered to be the main factor affecting wood viscoelasticity.  相似文献   

7.
The effect of cure history on the dynamic thermomechanical properties of a high temperature curing epoxy resin has been studied using torsional braid analysis. In isothermal cures “full cure” is not possible except at temperatures above the maximum glass transition temperature (Tg) of the cured resin, hence the necessity of a “post-cure” after lower temperature isothermal cures. The highest Tg and maximum cross-linking in the cured resin was for a linear heating rate of 0.05°C/min from 30 to 200°C; higher heating rates lead to lower glass transition temperatures.  相似文献   

8.
Raman spectroscopy of silicon nanostructures, recorded using an excitation laser power density of 1.0 kW/cm 2 , was employed here to reveal the dominance of thermal effects at temperatures higher than room temperature. The room temperature Raman spectrum showed only phonon confinement and Fano effects. Raman spectra recorded at higher temperatures showed an increase in FWHM and a decrease in asymmetry ratio with respect to its room temperature counterpart. Experimental Raman scattering data were analyzed successfully using theoretical Raman line-shapes generated by incorporating the temperature dependence of a phonon dispersion relation. The experimental and theoretical temperature dependent Raman spectra are in good agreement. Although quantum confinement and Fano effects persist, heating effects start dominating at temperatures higher than room temperature.  相似文献   

9.
A model N-methylol reactant system based on pyrrolidone derivatives has been used to study the chemical factors that control formaldehyde release. Fabric samples of known composition were prepared from chemically pure reactants and formaldehyde release was determined by the AATCC Sealed Jar Test. This work provides support for prior suggestions and a more direct approach for understanding unexpected results from tests on post-and precure fabrics. In addition, new findings demonstrate that the N-methylol reactant, its byproducts, and pure zinc nitrate are capable of reducing the level of formaldehyde release.  相似文献   

10.
To investigate the in vitro degradability and cytotoxicity of long alkyl cyanoacrylate polymers [polycyanoacrylates (PCAs)], we synthesized five kinds of alkyl cyanoacrylates (ethyl, 2‐octyl, n‐octyl, ethylhexyl, and ethyl cyanoacryloyllactate). In vitro degradation in buffer solutions and cell cultures for cytotoxicity were performed with PCAs prepared by various polymerization methods. Lower alkyl homologues such as ethyl cyanoacrylate revealed a higher tissue toxicity than higher alkyl homologues. The amounts of formaldehyde released from various PCAs were not proportional to the rate of degradation. The apparent form of the cyanoacrylate polymers greatly affected the degradation rate, as the powdery polymers degraded much more quickly than the films. A new biodegradable polymer, prepared from ethyl 2‐cyanoacryloyllactate, degraded more quickly than the others. The amount of formaldehyde released from the polymer degradation was high because it degraded rapidly. It was observed from cell culture experiments that the viability of the cells was higher with a lower release of formaldehyde because the alkyl side groups were bigger. Therefore, octyl cyanoacrylate polymers demonstrated lower amounts of formaldehyde by degradation and higher cell viability, and these monomers may be desirable for use as tissue adhesives. © 2003 Wiley Periodicals, Inc. J Appl Polym Sci 89: 3272–3278, 2003  相似文献   

11.
In order to develop zein fibers with improved physical properties and solvent resistance, formaldehyde was used as a crosslinking reagent before spinning. The crosslinking reaction was carried out in either acetic acid or ethanolic HCl where the amount of crosslinking reagent was between 1 and 6%. Reactions were carried out at various times and temperatures. When carried out in acetic acid, the maximum amount of formaldehyde that could be used was 1.5% as gelation occurred at higher levels. In ethanolic HCl, 6% formaldehyde could be incorporated into zein. All solutions were successfully electrospun, producing predominantly ribbon and round fibers; the morphology was strongly dependent on solvent and spinning solids. The formaldehyde‐crosslinked zein fabrics had double the tensile strength of control fabrics. SDS‐PAGE analysis clearly showed crosslinking had occurred. Electrospun fabrics from all formaldehyde‐treated zein solutions required an additional heating step in order to be resistant to dissolution in acetic acid, a known very good solvent for zein. Copyright © 2011 Society of Chemical Industry  相似文献   

12.
The curing behavior of synthesized phenol–urea–formaldehyde (PUF) resol resins with various formaldehyde/urea/phenol ratios was studied with differential scanning calorimetry (DSC) and dynamic mechanical analysis (DMA). The results indicated that the synthesis parameters, including the urea content, formaldehyde/phenol ratio, and pH value, had a combined effect on the curing behavior. The pH value played an important role in affecting the shape of the DSC curing curves, the activation energy, and the reaction rate constant. Depending on the pH value, one or two peaks could appear in the DSC curve. The activation energy was lower when pH was below 11. The reaction rate constant increased with an increase in the pH value at both low and high temperatures. The urea content and formaldehyde/phenol ratio had no significant influence on the activation energy and rate constant. DMA showed that both the gel point and tan δ peak temperature (Ttanδ) had the lowest values in the mid‐pH range for the PUF resins. A different trend was observed for the phenol–formaldehyde resin without the urea component. Instead, the gel point and Ttanδ decreased monotonically with an increase in the pH value. For the PUF resins, a high urea content or a low formaldehyde/phenol ratio resulted in a high gel point. The effect of the urea content on Ttanδ was bigger than that on the gel point because of the reversible reaction associated with the urea component. Too much formaldehyde could lead to more reversible reactions and a higher Ttanδ value. The effects of the synthesis conditions on the rigidity of the cured network were complex for the PUF resins. © 2005 Wiley Periodicals, Inc. J Appl Polym Sci 95: 1368–1375, 2005  相似文献   

13.
The changes in the hydrogen bonds (HBs) of three types of Indonesian lignite during low-temperature heating were investigated. The amount of water loss was determined by weighing the samples before and after heating in an oven. The changes in the number of the different types of HBs were determined using Fourier transform infrared spectroscopy coupled with types of water in lignite. The number of peak positions and absorption bands in each spectrum was determined by curve-fitting analysis with a Gaussian function. The quantified integrated area of aromatic hydrogen atoms was used to accurately investigate the changes in the HBs. The results show that at low temperatures (T ≤ 50°C), free water is mainly removed, and the HBs broken are those between free water molecules. However, at medium temperatures (50 <T ≤ 100°C), bound water is mainly removed. The number of HBs significantly changes because of the breaking of bound water molecule HBs and bound water cluster–carboxyl group HBs, and the formation of nonfreezable moisture HBs. At high temperatures (100 <T < 125°C), nonfreezable moisture can be released. The number of HBs changes as a result of competition between the removal of nonfreezable moisture and the increase in the number of carboxyl groups. At higher temperatures (T ≥ 125°C), the moisture remaining in lignite is thermal decomposition moisture. In addition, the rate of decomposition of carboxyl groups is higher than the rate of generation, which means that the number of HBs markedly decreases at higher temperatures.  相似文献   

14.
A series of amphoteric, superabsorbent polymers [poly(acrylic acid‐co‐diallyldimethylammonium chloride)] with different molar ratios of anionic groups to cationic groups were prepared by solution polymerization to investigate their swelling behaviors and the controlled release of agrochemicals. Various factors, including the solution pH, the concentrations of different salt solutions, and the temperature, were studied. The dynamic parameters of hydrogels at different temperatures suggested that diffusion was Fickian at lower temperatures, whereas non‐Fickian diffusion prevailed at higher temperatures. A copolymer hydrogel with a low anionic‐group/cationic‐group ratio showed a higher swelling capacity in water and higher salt tolerance. Also, the anionic‐group/cationic‐group ratio was not the dominant factor in determining the water retention. A poly(acrylic acid‐co‐diallyldimethylammonium chloride) hydrogel could control the release of agrochemicals effectively. © 2006 Wiley Periodicals, Inc. J Appl Polym Sci 102: 986–991, 2006  相似文献   

15.
Poly(N-isopropylacrylamide) gels exhibit a lower critical solution temperature (LCST) behavior in aqueous solution. At temperatures below the LCST, the gel is more hydrated than at temperatures above the LCST. At the LCST, the volume change is sharp. It is shown here that the water content of the gel above the LCST depends upon previous states of the gel (e.g., dry or wet), the heating rate, and the gel thickness. Deswelling kinetics are also affected by the gel thickness. Caffeine release experiments indicate that microscopic water pockets are formed within the matrix during the rapid gel collapse procedure. © 1994 John Wiley & Sons, Inc.  相似文献   

16.
Changes in the dynamic modulus Edyn of some phenolic glass cloth systems were studied as a function of cure temperatures and formaldehyde factors. The modulus changes were followed by continuously monitoring in situ the fundamental frequency of vibration of resin-impregnated glass cloth systems bonded to metallic cantilever reeds. The apparent activation energy of crosslinking was a direct function of the formaldehyde content in the resin synthesis, ranging from 9 kcal/mole for a 1.2 formaldehyde factor phenolic, to 20 kcal/mole for a 1.8 factor. Two cure parameters at 50% and 100% apparent mechanical conversion were established.  相似文献   

17.
Encapsulation of essential oils by in situ polymerization is commonly used to contain the oil and thus ensure its controlled release. Melamine resin formaldehyde is one of the most widely used shell materials due to its thermal and chemical stability. One of the factors that influences the properties of the capsules is the molar relationship between monomers. The effect of formaldehyde ? melamine (F/M) molar ratios 3, 4 and 6 on the nanoencapsulation, morphology and properties of nanocapsules was investigated. The morphology and particle size were investigated by the scanning electron microscopy and atomic force microscopy techniques. The composition of the formaldehyde ? melamine resins was determined by Fourier transform infrared spectroscopy, and the thermal stability of the nanocapsules was analysed by differential scanning calorimetry and thermogravimetric analysis. Increasing the formaldehyde content reduced the nanocapsules' chemical stability. The capsule sizes obtained were nanometric at all melamine ? formaldehyde ratios studied, with a non‐significant variation in particle size and shape. © 2017 Society of Chemical Industry  相似文献   

18.
The thermoelastic behavior for highly swollen films of syndiotactic-rich poly(vinyl alcohol) (s-PVA) was ascertained under loads in water through heating/cooling or cooling/heating cycles (in the range of 25–70°C). The s-PVA films kept at the high temperature of 70°C behaved as a perfect elastomer through the cooling/heating cycle and had very low Young's modulus, 2–3.5 × 106 dyn/cm2. When highly swollen s-PVA films were kept at lower temperatures below about 50°C for a long time, microcrystals were formed or propagated in the s-PVA films and plastic deformation occurred in addition to elastic deformation through heating process. The microcrystalline growth and propagation at lower temperatures were supported by an increase in Young's modulus and a decrease in the molecular weight between junctions.  相似文献   

19.
细木工板中甲醛释放特征及规律   总被引:2,自引:0,他引:2  
采用环境测试舱模拟室内环境,测量细木工板中甲醛的释放浓度,考察细木工板结构、温度、相对湿度和空气交换率对甲醛释放的影响,分析细木工板中甲醛气体扩散机理,并建立了灰色预测模型对细木工板中甲醛释放峰值后的过程进行模拟. 结果表明,细木工板中甲醛散发通道主要为端面,端面的甲醛释放量是平面的3倍;细木工板中甲醛气体扩散过程分为3个阶段(初始快速释放、稳定释放和长期缓速释放);空气交换率对细木工板中甲醛释放率影响不大;相对湿度和温度升高,细木工板中甲醛释放率也增大;预测模型的预测数据与实验数据吻合较好,平均相对误差率仅为3.717%,适合进行长期预测.  相似文献   

20.
p‐alkylphenol‐resorcinol‐formaldehyde‐latex (ARFL) films were prepared by co‐condensation of p‐alkylphenols and resorcinol with formaldehyde to generate modified phenolic resins, followed by blending with rubber latex, aging, and finally curing. The weight‐gain of the ARFL films and the tensile force of the coated fiberglass were studied under different temperatures and various humidities. The surfaces of the ARFL films were further analyzed by measuring the static contact angle and the findings were confirmed by Fourier Transform Infrared Spectroscopy (FTIR) and X‐ray photoelectron spectroscopy (XPS) analysis. The adhesion between the coated fiberglass and neoprene rubber was evaluated using the H‐adhesion technique. The best hydrophobicity and the largest water contact angle were displayed on the surface of the p‐nonylphenol‐resorcinol‐formaldehyde‐latex (NRFL) film, with a weight‐gain percent that was 40.0% (wt %) lower and a static contact angle that was 22.6° more than that of the resorcinol‐formaldehyde‐latex (RFL) film. The NRFL‐coated fiberglass had a higher tensile force and H‐adhesion force than the RFL‐coated fiberglass. The shelf life of NRFL‐coated fiberglass can be raised significantly at 40°C and under 98% humidity. The mechanism of the dramatic drop in the tensile force of the coated fiberglass is also discussed. © 2009 Wiley Periodicals, Inc. J Appl Polym Sci, 2009  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