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 ene reaction between plant oil triglycerides (such as soybean and sunflower oils) and paraformaldehyde was used to introduce a homoallylic hydroxyl functionality on the triglyceride. Paraformaldehyde and triglyceride were reacted in the presence of a Lewis acid catalyst, ethylaluminum dichloride, and hydroxymethyl derivatives were obtained at yields of 42 and 55% for sunflower oil and soybean oil, respectively. In the next step, hydroxymethyl products were reacted with maleic anhydride at 100°C to produce the maleate half esters. The average number of maleate groups per triglycerides was found to be 1.7 for soybean oil and 1.3 for sunflower oil. In the final step, the free‐radical–initiated copolymerization of the maleinized triglycerides with styrene produced rigid polymers. Characterization of new monomers and polymers was done by 1H‐NMR, 13C‐NMR, and infrared and mass spectrometries. The swelling behavior of the crosslinked network polymers was determined in different solvents. The glass‐transition temperature of the cured resin was also determined by differential scanning calorimetry to be 40°C for soybean‐based polymer and 30°C for sunflower‐based polymer. © 2004 Wiley Periodicals, Inc. J Appl Polym Sci 91: 4037–4046, 2004  相似文献   

2.
Epoxidized soybean oil (ESO) triglycerides were reacted with maleinized polybutadiene (MMPBD) to give plant‐oil‐based thermoset polymers. MMPBD samples were of two different molecular weights [high‐molecular‐weight maleinized polybutadiene (MMPBD‐H), maleate content = 10%, number‐average molecular weight (Mn) = 9000, and low‐molecular‐weight maleinized polybutadiene (MMPBD‐L), maleate content = 15%, Mn = 5000]. To increase the crosslink density of the product, a free‐radical initiator, benzoyl peroxide, was added to this mixture to further crosslink MMPBD through its double bonds. The characterizations of the products were done by dynamic mechanical analysis, differential scanning calorimetry, thermogravimetric analysis, and IR spectroscopy. The ESO–MMPBD polymers were crosslinked rigid infusible polymers. ESO–MMPBD‐H–1 : 1 and ESO–MMPBD‐L–1 : 1 showed glass‐transition temperature values at −23, 78 and −17, 64°C, respectively, whereas the storage moduli of the two polymers at 25°C were 13 and 16 MPa, respectively. The storage moduli of the polymers remained the same or decreased with the addition of a free‐radical initiator. The storage moduli also decreased with increasing ESO concentration above a 1 : 1 epoxy‐to‐anhydride molar ratio. The surface hardness increased dramatically, and the equilibrium swelling ratio decreased with the addition of free‐radical initiator. © 2010 Wiley Periodicals, Inc. J Appl Polym Sci, 2011  相似文献   

3.
Epoxidized soybean oil was reacted with cinnamic acid with triphenyl phosphine as a catalyst. Cinnamic acid reacted with 79% of the available epoxy groups, and this yielded cinnamate esters of epoxidized soybean oil (ESOCA). 1H‐NMR, IR, and mass spectra of the new cinnamate derivatives confirmed the proposed structure. The mass spectra revealed that the average number of cinnamate groups per triglyceride molecule was 3.33. ESOCA could be photopolymerized with UV light. ESOCA could also be homopolymerized into a soft and insoluble polymer by free‐radical initiation and copolymerized with styrene, vinyl acetate, and methyl methacrylate. A mixture of ESOCA with 25 wt % styrene had a viscosity of 410 cP and could be free‐radically polymerized with benzoyl tert‐butyl peroxide at elevated temperatures. Differential scanning calorimetry confirmed the formation of copolymers. The ESOCA homopolymer and its copolymers all showed a first‐order transition by differential scanning calorimetry around ?1.5°C that was attributable to side‐chain relaxations of the triglyceride fatty acids. The styrene copolymer of ESOCA showed a tan δ peak at 66.6°C. © 2003 Wiley Periodicals, Inc. J Appl Polym Sci 89: 3882–3888, 2003  相似文献   

4.
In this study, the polymerization reactions of epoxidized soybean oil (ESO) with the maleate half‐esters of oil‐soluble resoles and the properties of the final products were demonstrated. The maleate half‐esters of the dimeric oil‐soluble resoles were obtained by the esterification reaction of maleic anhydride with a p‐tertiary butyl phenol (p‐TBP) resole and p‐nonyl phenol resole resins in the first step. The monomers were characterized by IR and 1H‐NMR techniques. Then, the oil‐soluble resole maleates were polymerized with ESO to obtain tough and load‐bearing thermoset materials. The thermal and mechanical properties of the materials were determined by dynamic mechanical analysis, differential scanning calorimetry, thermogravimetric analysis, and tensile strength testing. The tensile strengths and storage moduli of the crosslinked polymers varied between 0.17 and 13 MPa and 10 and 1088 MPa, respectively. The elongation percentages of the materials were between 1 and 128%. The thermal resistance of the thermosets was measured as the 5% weight loss temperature. The reaction product of the ESO and maleate ester of p‐TBP showed the highest 5% weight loss at 247°C. © 2014 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2015 , 132, 41457.  相似文献   

5.
Maleinization followed by silylation of low molecular weight guayule rubber was accomplished for the first time via a two‐step process. Maleinization was first effected by grafting maleic anhydride onto guayule rubber under the influence of benzoyl peroxide catalyst to give maleinized guayule rubber (MAGR). The maleinized rubber was then further derivatized via hydrosilylation with dichloromethyl silane and chloroplatinic acid (Speier's catalyst) to give maleinized‐silylated guayule rubber (MASiGR). The silylated polymer (MASiGR) was further functionalized by reaction with ethanol or aqueous ammonia to produce ethoxysilane and silanol moieties, respectively. The products were characterized by Fourier transform infrared spectroscopy (FTIR), 1H‐, 13C‐, and 29Si‐ nuclear magnetic resonance spectroscopy (NMR). © 2001 John Wiley & Sons, Inc. J Appl Polym Sci 81: 754–761, 2001  相似文献   

6.
In this study, rigid thermoset polymers were prepared from radical copolymerization of linseed oil monoglyceride maleates with styrene (St). First, linseed oil monoglycerides (LOMGs) were obtained from the reaction of linseed oil with glycerol at 220–240°C. Then, LOMGs were reacted with maleic anhydride at 80°C to produce the LOMG maleate half esters. The reactions were followed by FTIR spectroscopy and size exclusion chromatography (SEC) and the final resin was characterized by 1H‐NMR spectroscopy. Finally, radical copolymerization of the LOMG maleates with 20, 40, 60, and 80% by weight of St was performed to produce rigid, thermoset polymers. The thermomechanical properties and fracture behavior of the cured copolymers, as a function of the percentage of St, were measured and analyzed. The copolymer with 40% by weight of St was the material with better mechanical and fracture behavior. © 2005 Wiley Periodicals, Inc. J Appl Polym Sci 97: 825–836, 2005  相似文献   

7.
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   

8.
In this study, styrene maleic anhydride copolymer (SMA2000, Styrene : Maleic Anhydride 2 : 1) is grafted and/or crosslinked with epoxidized methyl oleate, epoxidized soybean oil, methyl ricinoleate (MR), castor oil (CO), and soybean oil diglyceride. Base catalyzed epoxy‐anhydride and alcohol‐anhydride polyesters were synthesized by using the anhydride on SMA, the epoxy or secondary alcohol groups on the triglyceride based monomers. The characterizations of the products were done by DMA, TGA, and IR spectroscopy. SMA‐epoxidized soy oil and SMA‐CO polymers are crosslinked rigid infusible polymers. SMA‐epoxidized soy oil and SMA‐CO showed Tg's at 70 and 66°C, respectively. Dynamic moduli of the two polymers were 11.73 and 3.34 Mpa respectively. SMA‐epoxidized methyl oleate, poly[styrene‐co‐(maleic anhydride)]‐graft‐(methyl ricinoleate), and SMA‐soy oil diglyceride polymers were soluble and thermoplastic polymers and were characterized by TGA, GPC, DSC, NMR, and IR spectroscopy. © 2009 Wiley Periodicals, Inc. J Appl Polym Sci, 2010  相似文献   

9.
Soybean oil monoglycerides (SOMG), obtained by the glycerolysis of soybean oil, were reacted with maleic anhydride to produce SOMG maleate half esters. The copolymers of the SOMG maleates with styrene produced rigid thermosetting polymers. The dynamic mechanical analysis (DMA) of this polymer showed a glass‐transition temperature (Tg) around 133°C and a storage modulus (E′) value around 0.94 GPa at 35°C. The tensile tests performed on this polymer showed a tensile strength of 29.36 MPa and a tensile modulus of 0.84 GPa. Mixtures of SOMG with neopentyl glycol (NPG) and SOMG with bisphenol A (BPA) were also maleinized under the same reaction conditions and the resulting maleates were then copolymerized with styrene. The resulting polymers were analyzed for their mechanical properties. The Tg of the copolymers of the SOMG/NPG maleates with styrene was 145°C and the E′ value at 35°C was 2 GPa. The tensile strength of this polymer as calculated from the stress–strain data was 15.65 MPa and the tensile modulus was 1.49 GPa. The Tg of the copolymers of SOMG/BPA maleates, on the other hand, was found to be around 131°C and the E′ value was 1.34 GPa at 35°C. The changes observed in the mechanical properties of the resulting polymers with the introduction of NPG maleates and BPA maleates to the SOMG maleates can be explained by the structural changes on the polymer backbone. © 2002 Wiley Periodicals, Inc. J Appl Polym Sci 83: 972–980, 2002  相似文献   

10.
A novel thermal stabilizer poly(dioctyltin maleate‐styrene‐methyl acrylate) [P(DOTM‐St‐MA)] was synthesized by radical solution polymerization with benzene as the solvent and 2,2‐azobisisobutyronitrile as the initiator. The structure of terpolymer was characterized by FTIR and 1H‐NMR spectra, and thermal stability of the stabilizer was measured by thermogravimetric analysis (TGA). Evaluation of [P(DOTM‐St‐MA)] as thermal stabilizer for poly(vinyl chloride) (PVC) was measured by acidimeter, and the extent of changing color of PVC was measured by thermal aging method. Compatibilities of four stabilizers with PVC were characterized by scanning electron microscope (SEM). The results showed that, with the same tin content in PVC mixtures, [P(DOTM‐St‐MA)] exhibited better performance as a PVC stabilizer compared with other stabilizers,such as poly(dibutylin maleate‐styrene‐methyl acrylate), DOTM, and dibutylin maleate (DBTM). Furthermore, [P(DOTM‐St‐MA)] had better compatibility with PVC in PVC processing. © 2009 Wiley Periodicals, Inc. J Appl Polym Sci, 2009  相似文献   

11.
Epoxidized soybean oil (ESO) and isopropanolamine were used to synthesize a new polyol mixture for preparation of bio‐based polyurethanes. The chemical synthetic route for reaction of ESO with isopropanolamine was analyzed by 1H‐NMR. The results suggested that both ester groups and epoxy groups in ESO had reacted with amino group of isopropanolamine through simultaneous ring‐opening and amidation reactions. Epoxy groups in various situations exhibited different reactivity, and the unreacted epoxy groups were further opened by hydrochloric acid. The synthesized polyol mixture had high hydroxyl number of 317.0 mg KOH/g. A series of polyurethanes were prepared by curing the synthesized polyol mixture with 1,6‐diisocyanatohexance along with different amount of 1,3‐propanediol (PDO) as chain extender. Tensile tests showed that yield strengths of the polyurethanes ranged from 2.74 to 27.76 MPa depending on the content of PDO. Differential scanning calorimetry analysis displayed one glass transition temperature in the range of 24.4–28.7°C for all of the polyurethane samples, and one melt peak at high content of PDO. Thermogravimetric analysis showed that thermal degradations of the polyurethanes started at 240–255°C. In consideration of simple preparation process and renewable property of ESO, the bio‐based polyurethane would have wide range of applications. © 2012 Wiley Periodicals, Inc. J. Appl. Polym. Sci., 2013  相似文献   

12.
The Ritter reaction of plant oil triglycerides (such as soybean and sunflower oil) with acrylonitrile was used to introduce acrylamide functionality on the triglyceride. Acrylonitrile and triglycerides were reacted in the presence of H2SO4, and acrylamide derivatives were obtained in yields of 45 and 50% for sunflower oil and soybean oil, respectively. Radical initiated copolymerization of the acrylamide derivatives of the triglycerides with styrene produced semirigid polymers. Characterization of new monomers and polymers was done by 1H‐NMR, 13C‐NMR, IR, and MS. The swelling behavior of the crosslinked network polymers was determined in different solvents. Glass transiton temperature (Tg) of the cured resin was also determined by differential scanning calorimeter to be 40°C for soybean based polymer and 30°C for sunflower‐based polymer. Homo‐ and copolymerization behavior of acrylamide derivatives of methyl oleate (MOA) and methyl 10‐undecenoate (MUA) were also investigated. The reactivity ratios of these monomers with respect to styrene were determined by the Fineman–Ross method using 1H‐NMR spectroscopic data. The reactivity ratios were rsty = 1.776; rmoa = 0512 for MOA, and rsty = 1.142; rmua = 0.507 for MUA, respectively. Photopolymerization behaviors of MOA and MUA were also investigated using the photoDSC technique and the rate of polymerization of MUA is higher than that of MOA under the same conditions. © 2005 Wiley Periodicals, Inc. J Appl Polym Sci 97: 2264–2272, 2005  相似文献   

13.
Neslihan Alemdar  Niyazi Bicak 《Polymer》2010,51(22):5044-5050
Boric acid-pyridine mixture is presented as mild catalyst for polycondensation of semi-esters of ethylene glycol, in situ-generated by “cyclic anhydride-diol” reaction scheme from maleic, succinic and phthalic anhydrides. This catalyst system was demonstrated to give colorless polyesters in low molecular weights (Mn: 1650-1950 Da) within 4 h at 130 °C 1H NMR spectra of the polyesters derived from maleic anhydride revealed significant isomerization of maleate groups to fumarate units (∼80 %) during the polyesterification. Maleate and fumarate double bonds of the unsaturated polyesters (USP) were demonstrated to add bisulfite ions quantitatively to give sulfonated polyesters. The resulting sulfonated polyesters with 50% and 70% succinate units exhibited relatively narrow size spherical and ellipsoidal micelles in aqueous solutions, as inferred from surface tension, DLS and ESEM measurements. The results showed that, this procedure allows preparing a new family of polymer surfactants with tunable hydrophilicity and degradable polyester backbones.  相似文献   

14.
Four kinds of styrene/maleic–anhydride (SMA) copolymer‐derived dispersants with different charged forms were synthesized and characterized with 1H‐NMR. These four different dispersants contained carboxylic acid groups and exhibited characteristics indicative of anionic, cationic, zwitterionic, or nonionic surfactants as pH was changed. The charge properties of these copolymers changes with pH, and their ability to disperse clay particles under low and high pH were assessed by measuring viscosity or sedimentation, as well as Scanning Electron Microscopy (SEM). The results showed that the dispersing abilities are functions of pH of the system. It was found that SMA‐N [Poly(styrene‐co‐β‐N,N‐dimethylpropylamino) maleic acid, sodium salt)] exhibits the best dispersing ability, the fastest rate of sedimentation, and the smallest sedimentation volume at pH = 2, and SMA‐Na [poly(styrene‐co‐maleic acid, disodium salt)] exhibits the better dispersing ability at pH = 7 and 12. In addition, the aggregation behavior of the dispersants characterized by fluorescence spectroscopy revealed that the degree of aggregation for all dispersants at high concentration increases in the order: SMA‐Na < SMA‐B [poly(styrene‐co‐B‐imino‐propyl‐N‐trimethylammonium acetate) maleic acid, disodium salt)] < SMA‐N < SMA‐Q [poly(styrene‐co‐β‐imino‐propyl‐N‐trimethylammonium sulfate) maleic acid, sodium salt)]. © 2000 John Wiley & Sons, Inc. J Appl Polym Sci 78: 592–602, 2000  相似文献   

15.
Novel branched copolymers, poly(styrene‐alt‐maleic anhydride) (BPSMA), were synthesized through mercapto chain‐transfer polymerization with styrene, maleic anhydride (MA), and 4‐vinyl benzyl thiol (VBT). Then, the hydroxyl of hydroxyethyl methacrylate was reacted with MA to synthesize branched photosensitive copolymers, p‐BPSMAs. Fourier transform IR spectroscopy and 1H‐NMR indicated that the synthesis was successful. Gel permeation chromatography indicated that the molecular weight decreased with increasing content of VBT. The thermal properties were characterized by thermogravimetric analysis; the results show that the thermal decomposition temperature of the BPSMAs was greatly enhanced. Real‐time IR was used to evaluate the UV‐curable kinetics of the p‐BPSMAs; the results show that the p‐BPSMAs could rapidly photopolymerize under UV irradiation in the presence of photoinitiators. Moreover, the photoresist based on the p‐BPSMAs exhibited improved photosensitivity when the VBT content increased, and the photoresist with 12 mol % VBT content had a low value of the dose that retained 50% of the original film thickness (10 mJ/cm2), and a 50‐μm resolution could be achieved compared to a linear photoresist. © 2015 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2016 , 133, 42838.  相似文献   

16.
Maleated glycidyl 3‐pentadecenyl phenyl ether (M‐GPPE) was synthesized from glycidyl 3‐pentadecenyl phenyl ether (GPPE), a renewable derivative from cardanol, with maleic anhydride (MAH) by grafting copolymerization. The resulting M‐GPPE was used as a functionalized plasticizer for a styrene–butadiene rubber (SBR)/carbon black (CB)/silica composite. The effects of M‐GPPE on the development of the filler network, the extent of silica dispersion, the curing characteristics, and the mechanical performance of the composites were studied. Meanwhile, a comparative study was performed between M‐GPPE and aromatic oil, a traditional plasticizer used in SBR filler formulations. Gel permeation chromatography and IR and 1H‐NMR analysis results confirmed the occurrence of the grafting reaction between GPPE and MAH and the potential structure of M‐GPPE. The thermostability of GPPE was improved by grafting copolymerization with MAH, as shown by thermogravimetric analysis results. The presence of M‐GPPE resulted in a shorter curing time and better aging properties in the SBR composite compared with GPPE. The mechanical properties, dynamic mechanical analysis, and transmission electron microscopy analysis showed that the maleate of GPPE could enhance the compatibility between SBR and silica, improve the dispersion of silica in SBR, and partially replace the aromatic oil in the SBR/CB/silica composite formulation. © 2014 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2014 , 131, 40462.  相似文献   

17.
The grafting copolymerization of maleic anhydride (MAH) onto styrene‐butadiene‐styrene terpolymer (SBS) was carried out through a new synthesis method––solvothermal synthesis. Infrared (IR) spectra and solid state 13C‐NMR confirmed that maleic anhydride was successfully grafted onto the SBS backbone. The effects of different solvents, different initiators and their concentration, the amount of MAH, SBS concentration, and reaction time on the graft degree were evaluated, and the optimal conditions were obtained. Results indicated that the grafting reaction of MAH onto SBS through solvothermal method can be carried out in both good solvents and poor solvents, which are much different from the traditional solution grafting method, and high grafting degree can be obtained in good solvents. Finally, we also compared the grafting degree (GD) prepared by the solvothermal method with that by the melt grafting method and solution grafting method. © 2006 Wiley Periodicals, Inc. J Appl Polym Sci 102: 5274–5279, 2006  相似文献   

18.
Summary The configuration of maleic anhydride units in p-methoxystyrene-maleic anhydride copolymers prepared in methyl ethyl ketone at 50°C was studied using 13C NMR DEPT experiments. The ratio of cis (erythro) to trans (threo) configurations was found to increase with the tendency of the monomer units to alternate, remaining almost constant at approximately 1.33 when the mole fraction of maleic anhydride in the feed was larger than 0.30 and the monomer unit sequence was almost completely alternating. Also, para substitution on the styrene monomer units was found to greatly improve the resolution of the 13C NMR DEPT spectra for copolymers of p-methoxystyrene and p-chlorostyrene with maleic anhydride as compared with that of styrene-maleic anhydride copolymers.  相似文献   

19.
In this study 4‐methylpyridine (4MP), 4‐vinylpyridine (4VP) and poly(4‐vinylpyridine) (P4VP) were separately reacted with epoxidized soybean oil triglycerides (ESO) to give plant oil based thermoset polymers. The addition reaction of pyridine with epoxide followed by a rearrangement results in formation of pyridone units and these were polymerized via a Diels–Alder reaction. DMA, DSC, TGA and IR spectroscopy were used for the characterization of the products. 4MP‐ESO, P4VP‐ESO and P4VP‐ESO‐in situ polymers were crosslinked yielding rigid infusible polymers. Glass transition temperatures (Tg) of 4MP‐ESO and P4VP‐ESO‐in situ were found as ?10.5 and 70.5 (32.3 as shoulder) °C respectively, by DMA analysis. Storage moduli of 4MP‐ESO and P4VP‐ESO‐in situ at 25°C were 13.7 and 187.2 MPa, respectively. © 2011 Wiley Periodicals, Inc. J Appl Polym Sci, 2011  相似文献   

20.
A maleic anhydride grafted propylene–butadiene copolymer (MPPB) was prepared. Fourier transform infrared spectroscopy and 1H‐NMR results indicate that the maleic anhydride molecules reacted with the double bond in the butadiene unit of the propylene–butadiene copolymer (PPB), and the grafting percentage increased with the butadiene content in the initial copolymer. The gel permeation chromatography results show that the introduction of butadiene in the copolymer prevented the degradation of PPB. The MPPB was applied in polypropylene (PP)/styrene‐butadiene‐styrene triblock copolymer (SBS)/organophilic montmorillonite (OMMT) composites as a compatibilizer. In the presence of 10‐phr MPPB, the impact strength of the composite was improved by about 20%. X‐ray diffraction patterns indicated the formation of the β‐phase crystallization of PP in the presence of MPPB, and a significant decrease in the spherulite size was observed. Transmission electron microscopy (TEM) images showed that the OMMT was better dispersed in the matrix upon the inclusion of MPPB. A better distribution of the rubber phase and a rugged fracture surface were observed in the scanning electron microscopy images as the MPPB proportion was increased. © 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