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.
The structure of ZnO-catalyzed phenol–urea–formaldehyde (PUF) resin at different synthesis stages was analyzed by liquid 13C nuclear magnetic resonance spectroscopy. The results showed that the general structure of ZnO-catalyzed PUF resin was almost the same as the control PUF resin. Addition reaction between phenol and formaldehyde mainly occurred at the first stage. Total methylol groups amount between phenols of the control resin was a little lower than that of the ZnO-catalyzed PUF resin. Co-condensation and self-condensation reaction occurred at the second stage. The preparation method of ZnO-catalyzed PUF resin favored the co-condensation reaction between phenol methylol groups and urea units, while self-condensation reaction dominated the control resin at the second stage. Formaldehyde completely reacted for both the control and ZnO-catalyzed PUF final resin. The total amount of methylol and methylene groups between urea units and phenols, respectively, was almost the same for the two final resins. The total quantity of methylol groups between phenols represented a continuing downward trend from the first stage to the final stage, and the amount of methylol group (p-Ph–CH2OH) of ZnO-catalyzed PUF resin was 30% more than that of the control resin. Total co-condensed methylene groups amount of ZnO-catalyzed PUF resin was greater than that of the control resin, which indicated that ZnO could make the urea units well incorporated into the co-condensed PUF resin.  相似文献   

2.
红外光谱法对苯基苯酚改性酚醛树脂的研究   总被引:7,自引:0,他引:7  
本文利用红外光谱对苯基苯酚改性酚醛树脂进行了结构鉴定,并通过分析谱图中羟甲基、醚键含量的变化,研究了缩聚温度、聚合度对树脂结构的影响,结果表明苯基苯酚改性酚醛树脂从结构上就有利于获得较高的残碳率,而选择适当的反应条件可进一步提高树脂残碳率。  相似文献   

3.
The paper reports 13C nuclear magnetic resonance spectra of novolac resins and the variation of the composition of the reaction mixture during the in situ preparation of resin. It is observed that in the initial stages of the reaction ortho substitution occurs; however the hemiacetal rather than the methylol derivative is observed. These methylol substituted intermediates are highly unstable and a better appreciation of the relative rates of structure formation can be obtained from an analysis of the methylene bridge region, para-para Bridges are the first to appear, the next are the ortho-para and finally the ortho-ortho linkages are observed. Analysis of the relative intensities of ortho and para bridge carbon atoms allows identification of the isomeric composition of the final resin. The proportion of each isomer depends on the catalyst used in synthesis of the resin.  相似文献   

4.
Phenolic resins bearing varying concentrations of phenyl maleimide functions were synthesized by copolymerizing phenol with N‐(4‐hydroxyphenyl)maleimide (HPM) and formaldehyde in the presence of an acid catalyst. The resins underwent a two‐stage curing, through condensation of methylol groups and addition polymerization of maleimide groups. The cure characterization of the resin by dynamic mechanical analysis confirmed the two‐stage cure and the dominance of maleimide polymerization over methylol condensation in the network buildup process. The kinetics of both cure reactions, studied by the Rogers method, substantiated the earlier proposed cure mechanism for each stage. Although the initial decomposition temperature of the cured resin was not significantly improved, enhancing the crosslink density through HPM improved thermal stability of the material in a higher temperature regime. The anaerobic char yield also increased proportional to the maleimide content. Isothermal pyrolysis and analysis of the char confirmed that pyrolysis occurs by loss of hydrocarbon and nitrogenous products. The resins serve as effective matrices in silica‐ and glass fabric–reinforced composites whose mechanical properties are optimum for moderately crosslinked resins, in which failure occurs through a combination of fiber debonding and resin fracture. © 2001 John Wiley & Sons, Inc. J Appl Polym Sci 80: 1664–1674, 2001  相似文献   

5.
Gel formation characteristic of dilute phenol‐formaldehyde solution in water is studied at an elevated temperature near 140°C. It is experimentally found that at a temperature range of 135°C to 140°C, a solution containing 3.0% phenol with the phenol‐formaldehyde mole ratio 1.0:3.5 forms a gel in the pH range of 9.6 to 12.0. The minimum gelling time is observed at a pH value of 10.4. The gel formation is found to be strongly dependent on the pH at this elevated temperature. An explanation for the gelling behaviour is presented, which is based on an optimum ion concentration of Na+ to initiate the tri‐methylol formation. It is hypothesized that the gel formation begins when concentration of tri‐methylol phenol attains a threshold concentration value. Experimental observations and a kinetic analysis provide support for this hypothesis.  相似文献   

6.
二环戊二烯—苯酚树脂的合成及其在油性漆中的应用   总被引:8,自引:0,他引:8  
在三氟化硼催化剂存在下,采用二环戊二烯和苯酚合成了二环戊二烯-苯酚树脂,并用此树脂和亚麻油配制成油性漆。列举了漆膜的各项性能指标。讨论了不同摩尔比二环戊二烯和苯酚,以及反应温度对树脂的软化点和游离酚含量的影响。试验表明,热塑性二环戊二烯-苯酚树脂可作为涂料用树脂,在一定程度上可改善漆膜的气干性。  相似文献   

7.
周大鹏  范宏  卜志扬  李伯耿 《化工学报》2005,56(10):1987-1992
采用滴加苯酚的方式合成了N-苯基马来酰亚胺改性酚醛树脂(PPMF), 考察了N-苯基马来酰亚胺(PMI)参与酚醛缩聚反应的转化程度和树脂黏度特性的变化规律.发现延长苯酚滴加时间能提高PMI参与共缩聚反应的竞争能力和转化率;随着回流反应时间的增加,PMI转化率增加,树脂的黏度也随之增大;较高的反应温度或较大的PMI用量均会导致PMI转化率的显著下降.利用核磁、红外和元素分析等表征了改性酚醛树脂的结构及组成,确定N-苯基马来酰亚胺已共缩合于酚醛树脂中,其中PMI含量可达33%左右.利用DSC和TGA研究了N-苯基马来酰亚胺改性酚醛树脂的固化特性和耐热性能,发现PPMF树脂经历了两个固化阶段, 第一阶段是少量羟甲基的缩合, 第二阶段为马来酰亚胺的双键打开自交联形成. PPMF树脂固化产物的耐热性能优于传统的热塑性酚醛树脂.  相似文献   

8.
The chemical structure and cure characteristics of a group of phenolic resole resins were studied by means of three major analytical techniques. In particular, the effects on structure and reactivity of formaldehyde/phenol ratio and the type of reaction catalyst used were studied. Gel permeation chromatography was used to determine resin molecular weight distributions, and NMR, to determine chemical structural features. In this connection a selective oxidation procedure, converting free methylol groups to adehydes, has allowed unambiguous determination of methylene ether bridge structures to be made from the NMR data. The F/P ratio in a resole largely determines the type of molecular structures which are formed. However, triethylamine as a catalysts tends to favor methylene ether bridge formation, whereas sodium hydroxide favors methylene bridges. The rate and direction of subsequent thermal cure of the resoles prepared is shown by differential scanning calorimetry to depend markedly on the type of catalyst present during the curing stage. The DSC curing curves are interpreted in the light of the structural information provided by NMR.  相似文献   

9.
Some oligomeric species (condensates) in a melamine-formaldehyde adduct mixture have been isolated by preparative high performance liquid chromatography (h.p.l.c.) and partly identified by high field 1H and 13C nuclear magnetic resonance (n.m.r.) spectroscopy. The oligomers are shown most likely to be largely dimers, trimers and tetramers and to contain both methylene and methylene ether linkages. Some cured melamine-formaldehyde resins are shown by solid-state cross-polarisation magic angle sample spinning (c.p./m.a.s.) 13C and 15N n.m.r. to contain linear and branched methylene and methylene ether linkages but few free methylol groups. The relative proportions of the various linkages in both the oligomers and the cured resins appear to be controlled mainly by the melamine: formaldehyde ratio.  相似文献   

10.
The favored pH ranges for the formation of urons in urea‐formaldehyde (UF) resins preparation were determined, these being at pH's higher than 6 and lower than 4 at which the equilibrium urons ↔ N,N′‐dimethylol ureas are shifted in favor of the cyclic uron species. Shifting the pH slowly during the preparation from one favorable range to the other causes shift in the equilibrium and formation of a majority of methylol ureas species, whereas a rapid change in pH does not cause this to any great extent. UF resins in which uron constituted as much as 60% of the resin were prepared and the procedure to maximize the proportion of uron present at the end of the reaction is described. Uron was found to be present in these resins also as linked by methylene bridges to urea and other urons and also as methylol urons, the reactivity of the methylol group of this latter having been shown to be much lower than that of the same group in methylol ureas. Thermomechanical analysis (TMA) tests and tests on wood particleboard prepared with uron resins to which relatively small proportions of urea were added at the end of the reaction were capable of gelling and yielding bonds of considerable strength. Equally, mixing a uron‐rich resin with a low F/U molar ratio UF resin yielded resins of greater strength than a simple UF of corresponding molar ratio indicating that UF resins of lower formaldehyde emission with still acceptable strength could be prepared with these resins. © 1999 John Wiley & Sons, Inc. J Appl Polym Sci 72: 277–289, 1999  相似文献   

11.
A powdery product was obtained by the reaction of methylolated melamine with alkyl resorcinols to form melamine‐bridged alkyl resorcinols (MARs). The effects of the addition of this powder on the bonding strength and formaldehyde emission of urea–formaldehyde (UF) resins were investigated. Three types of UF resins with a formaldehyde/urea molar ratio of 1.3 synthesized by condensation at pH 1.0 (UF‐1.0), pH 4.5 (UF‐4.5), and pH 5.0 (UF‐5.0) were fabricated. The addition of MAR to UF‐4.5 and UF‐5.0 for bonding hardwood plywood enhanced the bonding strength and reduced formaldehyde emission. For UF‐1.0, the addition of MAR adversely affected the bonding strength. However, the UF‐1.0 resin yielded the lowest formaldehyde emission of all of the UF resins in the study. The effects of the MAR addition were related to the molecular structures of the UF resins. UF‐1.0 contained a large amount of free urea, a considerable number of urons, and a highly methylene‐linked, ring‐structured higher molecular weight fraction and had a smaller number of methylol groups. Therefore, the addition of MAR was considered to cause a shortage of the methylol groups, which in turn, led to incomplete resin curing. In contrast to UF‐1.0, UF‐5.0 contained a smaller amount of free urea and a linearly structured higher molecular weight fraction and had a larger number of methylol groups. In this case, MAR was considered to effectively react with the methylol groups to develop a three‐dimensional crosslinked polymer network to enhance the bonding strength and suppress the generation of free formaldehyde to reduce formaldehyde emission. © 2010 Wiley Periodicals, Inc. J Appl Polym Sci, 2010  相似文献   

12.
为降低耐火材料用酚醛树脂中游离酚的含量,选择苯酚与甲醛物质的量比1∶1.5,分步加入甲醛,醋酸锌为催化剂,添加质量分数2%,糠醇为溶剂,在60~85℃下制备了液态热塑性酚醛树脂,对产品的物化性能进行了测试并以红外光谱对其结构作了表征。结果表明,该工艺合成的酚醛树脂存在高邻位取代结构,残炭量达到45.66%,游离酚质量分数为1.87%。以原料中苯酚的量计算,产率可达到148.8%。  相似文献   

13.
Lot-to-lot variations in a phenol/cresol-formaldehyde resin were fully characterized. Resin property-product hardness relationships were examined to elucidate the cause of variable hardness in product insulators and to establish acceptance criteria for the resin. A high methylol content (?CH2OH), determined by quantitative carbon-13 nuclear magnetic resonance (13C-NMR), and a narrow molecular weight distribution (MWD), determined by size exclusion chromatography, were found necessary for resin to produce insulator meeting the minimum Shore D durometer specification of 40.  相似文献   

14.
采用两次碱一次回流法合成工艺代替一次法和两次碱法制备酚醛树脂胶黏剂,通过对酚醛比、反应时间、反应温度、回流时间等反应条件的考察得出:以n(苯酚):n(甲醛)=1:2的比例,在70℃下反应40 min,80℃下反应20 min后升温回流反应45 min得到具有交联结构的酚醛树脂胶黏剂,产品相对分子质量及黏度适中,拉伸剪切强度可达到5 MPa以上,游离甲醛含量低于0.03%,稳定性较好。  相似文献   

15.
Raw and dewaxed jute felt composites were prepared with resol and lignin modified phenol formaldehyde resin. Four different types of lignin modified resins were used by replacing phenol with lignin. The lignin modified resins were prepared from purified lignin obtained from paper industry waste black liquor. To investigate bonding between jute and resin, IR spectroscopy of jute felts and composites was carried out. The thermal stability of the composites was assessed by DSC and TGA. It was found that the lignin resin jute composite is thermally more stable than resol composite. XRD of jute felt and composite shows that the crystallinity of the jute fiber increases after composite preparation. The lignin resin composites were tested for water absorption and thickness swelling, and it was found that the results are comparable with those of resol jute composite. Composites prepared from lignin phenol formaldehyde resin with 50% phenol replacement has shown 75% tensile strength retention to that of pure resol jute composite.  相似文献   

16.
《Polymer》1986,27(5):750-756
Phenol-formaldehyde (PF) resins which had experienced a variety of reaction conditions and/or ageing conditions were quantitatively characterized by 13C n.m.r. Under the specific conditions of this study, the phenol para position was favoured for reaction over the ortho position on a per site basis, particularly for condensation reactions. Increasing the reaction temperature from 23°C (14 days reaction) to 80°C (3 h reaction) did not alter the type of resin structure. Ageing PF resins resulted in extensive condensation and a drastic reduction in para-substituted methylol and hemiformal groups. This apparently contributes to a sharp reduction in the reactivity of these resins for reactions requiring methylol substituents on the ring.  相似文献   

17.
A low‐formaldehyde‐emission methylol urea/triethanolamine composite was synthesized through in situ esterification of formaldehyde with triethanolamine and subsequent copolymerization of the synthesized polyester with methylol urea. The effects of the addition of triethanolamine to the polymerization process on some physical properties of the synthesized copolymer were evaluated. The copolymer was characterized with IR spectroscopy and macrophase‐separation techniques. At a given triethanolamine concentration, the composite exhibited macrophase‐separation behavior between that of pure methylol urea and pure polyester. IR spectra showed the presence of the polyester moiety in the composite. The values of the moisture uptake, formaldehyde emission, melting point, and elongation at break of the copolymer were within acceptable levels required in the coating industry. Therefore, the methylol urea/polyester copolymer resin could have potential as a binder in the coating industry. © 2009 Wiley Periodicals, Inc. J Appl Polym Sci, 2010  相似文献   

18.
研究了酚醛树脂合成条件对树脂中游离甲醛含量的影响。结果表明,反应物甲醛和苯酚的摩尔比为(1.2~2.4)∶1,树脂中游离甲醛的含量为0.12%~11.57%;将整个反应分成两步,在65℃和95℃各加热40 min,使游离甲醛含量更低;催化剂的最适合用量为苯酚摩尔数的6%~9%;游离甲醛含量随尿素用量增大而减小,但存在一个临界值,尿素的最佳用量为甲醛摩尔数的10%~12%。  相似文献   

19.
通过单因素实验,考察了影响制备高软化点热塑性酚醛树脂软化点、收率和树脂中游离酚的因素,并获得了优化的配方和工艺条件.结果表明,在苯酚与甲醛的物质的量之比为1∶0.9、催化剂为草酸、催化剂用量为1%(按苯酚的质量计算)、反应温度为90℃、反应时间为6h的条件下,所制备的线性酚醛树脂为无色透明、光泽度好的固体,软化点为12...  相似文献   

20.
Low‐condensation phenol‐formaldehyde (PF) resins coreacted under alkaline conditions with up to 42% molar urea on phenol during resin preparation yielded PUF resins capable of faster hardening times than equivalent pure PF resins prepared under identical conditions and presented better performance than the latter. The water resistance of the PUF resins prepared seemed comparable to pure PF resins when used as adhesives for wood particleboard. Part of the urea was found by 13C‐NMR to be copolymerized to yield the alkaline PUF resin; whereas, especially at the higher levels of urea addition, unreacted urea was still present in the resin. Increase of the initial formaldehyde to phenol molar ratio decreased considerably the proportion of unreacted urea and increased the proportion of PUF resin. A coreaction scheme of phenolic and aminoplastic methylol groups with reactive phenol and urea sites based on previous model compounds work has been proposed, copolymerized urea functioning as a prebranching molecule in the forming, hardened resin network. The PUF resins prepared were capable of further noticeable curing acceleration by addition of ester accelerators; namely, glycerol triacetate (triacetin), to reach gel times as fast as those characteristic of catalyzed aminoplastic resins, but at wet strength values characteristic of exterior PF resins. Synergy between the relative amounts of copolymerized urea and ester accelerator was very noticeable at the lower levels of the two parameters, but this effect decreased in intensity toward the higher percentages of urea and triacetin. 13C‐NMR assignements of the relevant peaks of the PUF resins are reported and compared with what has been reported in the literature for mixed, coreacted model compounds and pure PF and urea‐formaldehyde (UF) resins. The relative performance of the different PUF resins prepared was checked under different conditions by thermomechanical analysis (TMA) and by preparation of wood particleboard, and the capability of the accelerated PUF resins to achieve press times as fast as those of aminoplastic (UF and others) resins was confirmed. © 1999 John Wiley & Sons, Inc. J Appl Polym Sci 74: 359–378, 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