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 0 毫秒
1.
In this study, an easy and efficient synthesis of unsaturated plant oil triglycerides having isocyanate groups is reported. In the first step of the synthesis, the triglyceride was brominated at the allylic positions by a reaction with N‐bromosuccinimide, and in the second step, these brominated species were reacted with AgNCO to convert them to isocyanate‐containing triglycerides. At the end of the reaction, approximately 60–70% of the bromine was replaced by NCO groups, and the double bonds of the triglyceride were not consumed. When the amount of AgNCO was increased, the yield also increased. The final products were characterized with IR and 1H‐NMR, and polyurethanes and polyureas were obtained from these fatty isocyanates with alcohols and amines, respectively. The polymers were characterized by differential scanning calorimetry and thermogravimetric analysis. Differential scanning calorimetry curves showed that glycerin polyurethane showed a glass‐transition temperature at 19°C, castor oil polyurethane showed two glass‐transition temperatures at ?43 and 36°C, and triethylene tetraamine polyurea showed a glass‐transition temperature at 31°C. Some properties of the polymers, such as the tensile strength and swelling ratios, were also determined. The swelling rate of glycerin polyurethane was higher than that of castor oil polyurethane in dichloromethane. The equilibrium swelling ratio was highest for the castor oil polyurethane. The polyurethanes synthesized in this study had a Young's modulus around 50 kPa and a tensile strength around 0.01 N/mm2 (100 kPa). The tensile strength of glycerin polyurethane was higher than that of castor oil polyurethane. © 2008 Wiley Periodicals, Inc. J Appl Polym Sci, 2008  相似文献   

2.
Volatile by-products during heat polymerization of soybean oil at 330°C were analyzed using GC-MS and NMR. Color and viscosity changes were monitored for the heat-polymerized soybean oil and the by-products. About 90% (w/w) of the by-products were decanoic, palmitic, linoleic, oleic, and stearic acids and cis-9-tricosene. The by-products also contained small amounts of 3-eicosene, 9,17-octadecadienal, and cyclotetracosene. The weight percentage of decanoic acid increased with reaction time, whereas those of other components showed no trends.  相似文献   

3.
BACKGROUND: Vegetable oil‐based polyols are a new class of renewable materials. The structure of oil‐based polyols is very different from that of petrochemical polyols, and it is closely related to the structure of oils. The objective of this work was to analyze the structural heterogeneity of soy‐based polyols and its effect on the properties of polyols and polyurethanes. RESULTS: A series of polyols with a range of hydroxyl numbers were prepared by hydroformylation and partial esterification of hydroxyls with formic acid. Polyols were reacted with diphenylmethane diisocyanate to obtain polyurethanes of different crosslinking density. Gelation was simulated using the Monte Carlo method with a calculated distribution of functionalities for each polyol. CONCLUSIONS: Most polyols are powerful crosslinkers since weight average functionality varied from 5 to 2.5 resulting in gel points from 53 to 83% conversion. Heterogeneity of polyols had a negative effect on mechanical properties of rubbery polyurethanes and this should be taken in account when designing polyols for flexible applications. This effect was not pronounced in glassy polyurethanes. Copyright © 2007 Society of Chemical Industry  相似文献   

4.
紫外光固化环氧豆油丙烯酸酯的制备与表征   总被引:3,自引:0,他引:3  
将环氧大豆油与丙烯酸反应制备出环氧豆油丙烯酸酯预聚物,讨论了反应温度、反应时间、催化剂、阻聚剂的种类与用量对合成反应的影响,并用红外光谱对产物的结构进行了表征。研究结果表明其最佳反应条件是:催化剂三苯基膦,反应温度110℃,反应时间8h。该预聚树脂可用紫外光固化,其固化膜硬度达3H,且具有较好的柔韧性和附着力。  相似文献   

5.
Soybean oil contains five triglycerides that account for over 70% of the total. These include LLL, LLO, LLP, LOO, and LOP and their isomers (where L=linoleic, O=oleic, P=palmitic). High-performance liquid chromatographic analysis of samples taken during a typical hydrogenation run in which the iodine value (IV) was reduced from 130 to about 70 showed that the linoleate-containing triglycerides were reduced at a much faster rate than the linolenate-containing triglycerides. Results clearly show that hydrogenation proceeds through definite pathways rather than by random saturation of fatty acids within the triglyceride molecules. Pressure has a significant effect on the course of hydrogenation. At higher pressures (500 psi), the reaction is truly nonselective, since di-and trisaturated glycerides are formed at about 70 IV, whereas at 50 psi, the reaction becomes selective. At higher pressures, fat functionality and solid fat result primarily from di-and trisaturated fatt acid triglycerides as well as trans fatty acid triglycerides. At lower pressures, functionality results from trans acid triglyceride formation. Although the reactivity of linoleate containing triglycerides followed the pattern 6 double bonds>5 double bonds>4 double bonds, other factors may be important. For example, LLP is reactive and undergoes hydrogenation, while LLS remains unchanged. Triolein, which constitutes less than 3% of the total triglyceride in natural soybean oil, is a significant product of hydrogenation, which suggests that LLL and LLO are reduced directly while adsorbed on the catalyst surface. Presented in part at the A.E. Bailey Award Dinner, Peoria, IL, Feb. 23, 1999.  相似文献   

6.
Rigid polyurethane foam (RPUF) is prepared from petroleum-based polyols and isocyanate, which consumes a large amount of petroleum. To alleviate the consumption of petroleum, it is necessary to synthesize green and sustainable polyols. However, the greatest disadvantage of RPUF is its flammability. To reduce the risk of fire caused by RPUF, phosphorylated soybean oil polyol (Polyol-P) and phenyl phospho-soybean oil polyol (Polyol-PPOA) were synthesized by ring-opening reactions of epoxy soybean oil with phosphoric acid and phenylphosphonic acid, respectively. A flame-retardant RPUF was prepared via polymeric 4,4-diphenylmethane diisocyanate (p-MDI), which reacted after mixing Polyol-P and Polyol-PPOA with polyether polyol-330N in different proportions. Scanning electron microscopy (SEM) showed that the cell sizes of the RPUF-P and RPUF-PPOA increased first and then decreased and the cell number decreased first and then increased with the increase in the contents of Polyol-P and Polyol-PPOA. Mechanical property tests showed that the compressive strength of the RPUF-P4 reached 0.1 MPa, and the compressive strength of the RPUF-PPOA4 reached 0.07 MPa. The limiting oxygen index values of the RPUF-P4 and RPUF-PPOA4 were 20.9% and 24.3%, respectively. The UL 94 of the RPUFs indicated that the rating of the RPUF-PPOA3 was improved to V-1. The results showed that the flame-retardancy mechanism of the Polyol-P and Polyol-PPOA in the RPUF was based on the charred surface as a physical barrier, which slowed down the decomposition of RPUF and prevented heat and mass transfer between the gas and the condensed phase.  相似文献   

7.
Both rigid and flexible water-blown polyurethane foams were made by replacing 0–50% of Voranol® 490 for rigid foams and Voranol® 4701 for flexible foams in the B-side of foam formulation by epoxidized soybean oil. For rigid water-blown polyurethane foams, density, compressive strength, and thermal conductivity were measured. Although there were no significant changes in density, compressive strength decreased and thermal conductivity decreased first and then increased with increasing epoxidized soybean oil. For flexible water-blown polyurethane foams, density, 50% compression force deflection, 50% constant force deflection, and resilience of foams were measured. Density decreased first and then increased, no changes in 50% compression force deflection first and then increased, increasing 50% constant force deflection, and decreasing resilience with increase in epoxidized soybean oil. It appears that up to 20% of Voranol® 490 could be replaced by epoxidized soybean oil in rigid polyurethane foams. When replacing up to 20% of Voranol® 4701 by epoxidized soybean oil in flexible polyurethane foams, density and 50% compression deflection properties were similar or better than control, but resilience and 50% constant deflection compression properties were inferior. © 2008 Wiley Periodicals, Inc. J Appl Polym Sci, 2008  相似文献   

8.
Methyl oleate, a kind of vegetable oil derivatives, was used as the raw material to synthesize novel polyols containing nitrogen and phosphorus (PNP) by three steps: amidaion, epoxidation, and ring-opening reaction. The structure of the ensuing PNP was characterized by gel permeation chromatography, Fourier transform infrared spectroscopy, and nuclear magnetic resonance spectroscopy. Then, polyurethane films were prepared by mixing the resulting polyols (PNP) and poly-(tetra methylene glycol) with isophorone diisocyanate in different proportions. After that, the thermostability of polyurethane films was investigated by limiting oxygen index (LOI), thermogravimetry, and cone calorimeter test. Similarly, mechanical property of polyurethane films was studied by tensile test. The results showed that as the amount of PNP increased, the tensile strength of polyurethane films increased from 2.0 to 13.5 MPa, the LOI values of polyurethane films went up from 20.1% to 25.5%. Moreover, the initial decomposition temperature of polyurethane films improved from 280 to 312°C and the total heat release drooped from 32.2 to 21.7 kW/m2 without adding any other flame retardant. Therefore, it was observed that the polyurethane films have superior thermal stability and mechanical property and appear suitable for a wide range of applications.  相似文献   

9.
Grubbs' ruthenium catalyst 2 has been employed in model studies of the acyclic diene metathesis (ADMET) polymerization of soybean oil. In the presence of 0.1 mol% of catalyst 2, the ADMET polymerization of ethylene glycol dioleate afforded the isomerized (E)-dioleate (27%), dimer (18%), trimer (13%), tetramer (7%), pentamer (5%), hexamer (4%), heptamer (4%), and 9-octadecene (21%). Only a trace of any intramolecular cyclized product was formed. Under the same conditions, glycerol trioleate underwent ADMET polymerization to produce dimer, trimer, tetramer, pentamer, and monocyclic oligomers, with monocyclic oligomers predominating. The high number of repeat units in the monocyclic oligomers (n≅6, 10, and 21) in dicates that cross-linking occurs readily in this process. Based on our model system studies, we have examined the ADMET polymerization of soybean oil and succeeded in producing polymeric materials ranging from sticky oils to rubbers.  相似文献   

10.
Polyurethanes can be prepared using polyols obtained from vegetable oils in natura, such as castor oil, or from functionalized vegetable oils, such as hydroxylated soybean oil. These polyurethanes have different valuable properties, determined by their chemical composition and cross-linking density. In this study, soy epoxy polyols with different OH contents were prepared through a one-step reaction using the method of in situ performic acid generation. Polyols with OH functionalities from 1.9 to 3.2 were reacted in bulk with different diisocyanates at a NCO/OH molar ratio of 0.8 and 60°C for 24 h. Mechanical properties of the polyurethanes were determined by dynamic mechanical thermal analysis, hardness (Shore A), and swelling measurements. Polymer networks with glass-transition temperatures (T g ) from −13 to 48°C were obtained. We observed that the higher the OH functionality of the polyols, the higher the T g and cross-linking density of the polyurethane network. The influence of diisocyanate structure (rigid or flexible chain), curing temperature, and curing reaction time on mechanical properties was also investigated.  相似文献   

11.
Acrylated epoxidized soybean oils (AESOs) with different level of saturation were obtained by the ring opening of different saturation epoxidized soybean oils using acrylic acid as the ring opener. AESO‐based thermosets have been synthesized by free radical polymerization of these AESOs and methyl methacrylate. The thermal properties of these resins were studied by differential scanning calorimetry and thermo‐gravimetric analysis. The results indicated that the thermal stability of these resins depends upon the epoxy value; the glass transition temperature increases with increasing of epoxy value. The tensile and impact strength of the resins were also studied, and indicated that tensile strength increases with increasing epoxy value, whereas impact strength decreases. The resulting thermosets ranged from elastomers to glassy polymers.  相似文献   

12.
Zengshe Liu  Sevim Z. Erhan 《Polymer》2005,46(23):10119-10127
New epoxidized soybean oil (ESO)/clay nanocomposites have been prepared with triethylenetetramine (TETA) as a curing agent. The dispersion of the clay layers is investigated by X-ray diffraction (XRD) and transmission electron microscopy (TEM). XRD and TEM data reveal the intercalated structure of ESO/clay nanocomposites has been developed. The thermogravimetric analysis exhibits that the ESO/clay nanocomposites are thermally stable at temperatures lower than 180 °C, with the maximum weight loss rate after 325 °C. The glass transition temperature, Tg, about 7.5 °C measured by differential scanning calorimetry (DSC) and Tg about 20 °C measured by dynamic mechanical study have been obtained. The difference in the Tg between DSC and dynamic measurements may be caused by different heating rate. The nanocomposites with 5-10 wt% clay content possess storage modulus ranging from 2.0×106 to 2.70×106 Pa at 30 °C. The Young's modulus (E) of these materials varies from 1.20 to 3.64 MPa with clay content ranging from 0 to 10 wt%. The ratio of epoxy (ESO) to hydrogen (amino group of TETA) greatly affects dynamic and tensile mechanical properties. At higher amount of TETA, the nanocomposites exhibit stronger tensile and dynamic properties.  相似文献   

13.
In this study, a novel and simple route for the synthesis of the iodine isocyanate (INCO) adduct of soybean oil triglycerides is described. Soybean oil iodo isocyanate (ISONCO) was synthesized by the reaction of iodine isocyanate and soybean oil at room temperature. ISONCO was then polymerized with polyols, such as, castor oil, pentamethylene glycol, and glycerol to give the corresponding polyurethanes and with polyamines, such as, ethylene diamine, hexamethylene diamine, and triethylene tetramine to give corresponding polyureas. The structures of the monomer and the polymers were determined by FTIR and 1H‐NMR analyses. Thermal properties of the polymers were determined by DSC and TGA. Thermal degradation of the polyurethanes started at 150°C. Stability of the polyureas was higher than polyurethanes. Almost all polymers showed a Tg around ?50°C. The mechanical properties of the polymers were determined by tensile tests. Among the polymers synthesized, castor oil polyurethane showed the highest elongation at break and the lowest tensile strength of 140 KPa. The highest tensile strength of 900 KPa was observed in the pentamethylene glycol polyurethanes. © 2010 Wiley Periodicals, Inc. J Appl Polym Sci, 2010  相似文献   

14.
以天然大豆油为原料,采用酯基插入式乙氧基化技术一步法合成了天然大豆油乙氧基化物,并对产品生理毒性进行了测试。结果表明,在催化剂MCT-09作用下,大豆油与环氧乙烷可顺利合成不同EO数的大豆油乙氧基化物,与其他类型非离子表面活性剂相比,天然大豆油乙氧基化物生理毒性最小,为无毒类绿色表面活性剂,并通过红外光谱对产物进行了结构鉴定。  相似文献   

15.
In this study, rigid thermoset polymers were prepared from radical copolymerization of the soybean oil monoglyceride maleates with styrene. In the first part of the study, soybean oil monoglycerides (SOMGs) were obtained from the reaction of soybean oil with glycerol at 220–240°C with an optimization of the reaction to maximize the monoglyceride yield. In the following step, SOMG were reacted with maleic anhydride at temperatures around 100°C to produce the SOMG maleate half esters. Different catalysts and different reaction conditions were examined to increase the maleate half esters' yields. The reactions were followed by IR and 1H NMR, and the products were characterized by mass spectrometry. In the final step, the radical initiated copolymerization of the SOMG maleates with styrene produced rigid, thermoset polymers. The emulsion copolymerization of the SOMG maleates with styrene was also carried out successfully without the addition of an emulsifier. The obtained polymers were characterized by IR and the crosslinked network structure of the copolymers was examined with the swelling behavior in different solvents. Mechanical properties of the cured resin such as Tg, dynamic flexural modulus, and surface hardness were also determined. © 2001 John Wiley & Sons, Inc. J Appl Polym Sci 81: 69–77, 2001  相似文献   

16.
In this study, rubbery, bio‐based thermoset polymers were synthesized from soybean oil (SO) and p‐dinitrosobenzene (DNB) via an ene reaction. Polymeric p‐dinitrosobenzene (PDNB) was synthesized from p‐quinone dioxime and was thermally depolymerized in the presence of SO. SO/PDNB polymers were synthesized by a two‐stage polymerization. During the reaction, the role of different parameters such as mol ratio of SO and PDNB, preheating temperature of SO, polymerization time, and temperature were examined. Polymerization was followed by IR spectroscopy, and the polymers obtained were characterized by dynamic mechanical analysis, thermogravimetric analysis, and differential scanning calorimetry. The polymers have glass transition temperature ranging from ?51.5°C to ?46.3°C, whereas their storage moduli are between 430 and 210 MPa at ?60°C. Thermogravimetric analysis reveals that all of the polymers have temperatures of maximum degradation around 500°C. The crosslinked network structure of the polymers was investigated by swelling behavior and surface hardness test. Methyl oleate was used as a model compound to examine the chemical structure of the products. © 2009 Wiley Periodicals, Inc. J Appl Polym Sci, 2009  相似文献   

17.
Traditional plastics are usually obtained from nonrenewable petroleum feedstocks. In this study, crosslinked polymers were synthesized from cottonseed oil. Unsaturated fatty acids, the major components of cottonseed oils, were initially epoxidized. A network polymer was then formed by crosslinking the epoxidized oil with maleic anhydrate. Mechanical properties of these polymers were altered by varying the amount of maleic anhydrate. These polymers have a tensile modulus of the order 1 MPa and are stable in the acidic and alkaline environment. Our results provide a new synthetic strategy to obtain a network polymer from cottonseed oil. © 2019 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2019 , 136, 47655.  相似文献   

18.
A series of flexible polyurethane foams with different polyol compositions were synthesized through the replacement of a portion of the petroleum‐based polyether polyol with biobased polyols, namely, glycerol (GLY) and hydroxylated methyl esters (HMETO). HMETO was synthesized by the alkaline transesterification of tung oil (TO; obtaining GLY as a byproduct) and the subsequent hydroxylation of the obtained methyl esters with performic acid generated in situ. FTIR spectroscopy, 1H‐NMR, and different analytical procedures indicated that the hydroxyl content increased significantly and the molecular weight decreased with respect to those of the TO after the two reaction steps. The characterization of the obtained foams, achieved through the measurement of the characteristic reaction times, thermal and dynamic mechanical analysis, scanning electronic microscopy, and density measurements, is reported and discussed. The most important changes in the modified foams were found with the addition of GLY to the formulation; this led to an increased foam density and storage rubbery modulus, which were associated with a higher crosslinking density because of the decrease in the chain length between crosslinking points. © 2016 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2016 , 133, 43831.  相似文献   

19.
The cationic copolymerization of tung oil with the divinylbenzene comonomer initiated by boron trifluoride diethyl etherate produces promising plastics. The gel times are largely dependent on the relative composition and the reaction conditions and vary from a few seconds to 1 h. Controlled reactions producing homogeneous materials can be obtained by (1) lowering the reaction temperature or (2) decreasing the initiator concentration to less than 1 wt % or (3) adding a certain amount of a less reactive oil, such as soybean oil, low saturation soybean oil (LoSatSoy), or conjugated LoSatSoy to the reaction. The resulting polymers are rigid and dark brown in color. The weight % of the starting materials converted to the crosslinked polymer is ∼85–98% as determined by Soxhlet extraction with methylene chloride. The structure of the bulk product is that of a crosslinked polymer network plasticized by a small amount of low molecular weight oil. The chemical composition of the bulk polymers varies with the original composition of the tung oil system. Dynamic mechanical analysis shows that the resulting products are typical thermosetting polymers with densely crosslinked structures. The modulus of the plastics is approximately 2.0 × 109 Pa at room temperature. One broad glass transition is observed at approximately 100°C. Thermogravimetric analysis shows that the tung oil polymers are thermally stable below 200°C with a 10% weight loss in air around 430°C. © 2000 John Wiley & Sons, Inc. J Appl Polym Sci 78: 1044–1056, 2000  相似文献   

20.
用碱催化酯交换法合成大豆油异丙酯,考察了醇油摩尔比、催化剂用量、反应温度、反应时间对酯交换反应的影响。结果表明,合成大豆油异丙酯的最佳条件如下:醇油摩尔比10∶1,反应时间60 min,反应温度80℃,碱催化剂浓度2.0%。在此条件下,大豆油异丙酯的产率达53.07%。最后对比研究了大豆油甲酯、乙酯、异丙酯的合成,得出不同结构的醇与大豆油进行酯交换反应的规律。色谱表征表明,经酯交换作用,原料油成功转化为生物柴油。  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