首页 | 官方网站   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 thermal stability and degradation kinetics of several polystyrenes and styrene–alkyl methacrylate copolymers and terpolymers with a number-average molecular weight (M?n) of 6000–250,000 g/mole have been studied using dynamic thermogravimetry (TG). The degradation kinetics of each polymer sample have been successfully attributed to a sample first-order reaction expression. The results indicate that the thermal stability and degradation kinetics of the polymers are independent of the size of the molecules within the molecular weight range investigated. The steric hindrance effects of the pendent groups appear to be responsible for the improved thermal stability and resistance of C? C bond scission in the styrene–alkyl methacrylate copolymers and terpolymers.  相似文献   

2.
A styrene–butadiene–styrene (20–60–20) linear block polymer was epoxidized in toluene and cyclohexane solutions with peroxyformic acid generated in situ. The epoxidized polymer, containing about 8% oxygen, exhibits greatly improved resistance to ASTM oils. It can be readily compounded with carbon black, and the unvulcanized stock is found to be comparable to vulcanized polychloroprene and nitrile rubber in tensile strength and resistance to ASTM oils and certain chemicals.  相似文献   

3.
The degradation behavior of mixtures of polystyrene and poly(methyl methacrylate) in the form of thin films cast from a solution containing both polymers has been compared with that of the individual polymers and of copolymers of the same monomer pair by means of thermal volatilization analysis (TVA) together with product analysis by gas-liquid chromatography determinations of the molecular weight of the solid polymer residues. The results indicate no interaction between the polymers when degraded together. The polymer mixtures are readily distinguished from copolymers of the same overall composition by TVA.  相似文献   

4.
Epoxidation of styrene–butadiene block polymers considerably improves their resistance to hydrocarbon oils. Use of peroxyformic acid generated in situ appears to be simple and practical. Although no problems exist when the reactions are carried out in toluene, polymer molecular weight and reactant stoichiometry are important in determining the properties of polymers epoxidized in cyclohexane. In fact, epoxidations in cyclohexane of linear and tapered block polymers of molecular weights in excess of 60,000 with peroxyformic acid invariably lead to gelation. Replacing half or more of formic acid with acetic acid, however, alleviates the problem without adversely affecting the tensile or oil resistance of the resulting polymers. This recipe has also been successfully applied to the epoxidation of divinylbenzene-coupled styrene–butadiene radial block polymers.  相似文献   

5.
The glass transition temperatures of the poly(alkyl α-cyanocrylates) were determined by the dilatometric technique, and some of the values were checked by differential thermal analysis. The data indicate that the Tg's appear to decrease with increase in the size of the alkyl group, for a given molecular weight range. It was also found that the Tg of poly(methyl or butyl α-cyanoacrylate) increased with molecular weight. All cyanoacrylates, excepting methyl and ethyl esters, formed only low molecular weight polymers in aqueous surroundings. Therefore, they have characteristic low glass transition temperatures, causing coalescence at low temperatures.  相似文献   

6.
Thermal stability of PVC blends with chlororubber-20-graft polyblend-styrene-acrylonitrile [CR-20gp-SAN (2:1)] was studied by HCI evolution techniques and thermogravimetry under isothermal condition. The thermal stability of PVC/CR-20gp-SAN (2:1) blends has been compared with those of PVC/CR-20 and PVC/KM-365B blends. It has been observed that the thermal stability of modified PVC is less than that of unmodified PVC. The CR-20gp-SAN (2:1) modified PVC blends were found to be more stable than PVC/CR-20 blends but less stable than PVC/KM-365B blends. The rate of degradation in PVC blends were observed to be unaffected by the concentration of the modifiers, but the PVC/KM-365B blends were found to be degrading slower in comparison with PVC/CR-20 and PVC/CR-20gp-SAN (2:1) blends. The rate of degradation for PVC/CR-20 blends at lower concentrations (<10%) of modifiers is almost equal to that of PVC/CR-20-gp-SAN (2:1) blends, but more at higher concentrations of modifiers (>10%). The experimental results have been explained on the basis of the chemical nature of the modifiers and their miscibility with PVC.  相似文献   

7.
Compositional analysis of styrene–methyl methacrylate random copolymers by UV spectrophotometry at 260 nm, proton-NMR, and IR at 1730 cm?1 have been accomplished and agreement between these three independent methods was excellent. IR spectra of the copolymers in the region of 1100 and 1300 cm?1 are mainly characteristic absorption bands for the methyl methacrylate (MMA) component, but not the same for those of MMA homopolymer, that is, the IR spectra of the copolymers were not additive with those of polystyrene and PMMA. Information about the sequence distribution of copolymers of the same composition can be obtained by comparing the wavenumbers and absorption coefficients of the IR spectra in the region of 1100–1300 cm?1  相似文献   

8.
The thermal properties of ethylene–propylene copolymer grafted with glycidyl methacrylate (EP-g-GMA) were investigated by using differential scanning calorimetry (DSC). Compared to the plain ethylene–propylene copolymer (EP), peak values of melting temperature (Tm) of the propylene sequences in the grafted EP changed a little, crystallization temperature (Tc) increased about 8–12°C, and melting enthalpy (ΔHm) increased about 4–6 J/g. The isothermal and nonisothermal crystallization kinetics of grafted and ungrafted samples was carried out by DSC. Within the scope of the researched crystallization temperature, the Avrami exponent (n) of ungrafted sample is 1.6–1.8, and those of grafted samples are all above 2. The crystallization rates of propylene sequence in EP-g-GMA were faster than that in the plain EP and increased with increasing of grafted monomer content. It might be attributed to the results of rapid nucleation rate. © 1996 John Wiley & Sons, Inc.  相似文献   

9.
The prepolymers for a novel oil absorbent were synthesized by copolymerizing styrene with 2‐ethylhexyl acrylate (EHA), lauryl acrylate (LA), lauryl methacrylate (LMA), and stearyl acrylate (SA). Suspension polymerization was carried out using benzoyl peroxide (BPO) as an initiator with a varying monomer feed ratio, and the copolymers were characterized by FTIR, 1H‐NMR, DSC, and a solubility test. The copolymers were random copolymers with a single phase, and their compositions were similar to those in the monomer feed. The Tg of the copolymer could be controlled by varying the styrene/acrylate ratio. Acrylates introduced the crosslinking to linear polymers as a side reaction. Crosslinked copolymers were synthesized by adding divinylbenzene (DVB) as a crosslinking agent. At a low degree of crosslinking (0.5 wt % DVB), the Tg of the crosslinked copolymers was lower than or similar to that of the uncrosslinked ones. At a high degree of crosslinking, the Tg increased with increasing crosslinking density. © 2000 John Wiley & Sons, Inc. J Appl Polym Sci 77: 903–913, 2000  相似文献   

10.
Amphiphilic polymers have found many applications, so many types of these copolymers have been prepared. Specifically, sulfonated polystyrene acts, for example, as a flocullant or dispersant of petroleum asphaltenes as a function of its hydrophilic–hydrophobic balance. However, when changing the sulfonation degree, looking for the best performance, the solubility also changes, and sometimes it is responsible for making the polymer unsuitable for any application. Therefor, this work investigates in detail the changes in the solubility range of copolymers based on styrene–stearyl methacrylate and styrene–stearyl cinnamate with different molar compositions and different sulfonation degrees. The copolymers were synthesized and characterized by 1H‐NMR, Fourier transform infrared spectroscopy, and elemental analysis. In the range of compositions analyzed, with increasing content of long hydrocarbon chains, not only the displacement of the solubility in solvents with lower solubility parameter (δ), but also the broadening of the solubility range was observed. In general, the solubility was directly related to the sulfonic group content, but there appeared to be an influence of the randomness of the sulfonation reactions along the chains. © 2015 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2016 , 133, 43112.  相似文献   

11.
Multiblock copolymers of styrene, butadiene, and ethylene oxide were synthesized by coupling together telechelic dihydroxyl polystyrene, telechelic dihydroxyl polybutadiene, and poly(ethylene glycol), using 2,4-toluene diisocyanate as coupling agent. The copolymers were purified by extractions and characterized by infrared (IR), 1H nuclear magnetic resonace (NMR), gel permeation chromatography (GPC), transmission electron microscopy (TEM), membrane osmometry, and dynamic viscoelastometry. The multiblock copolymers are amphiphilic, exhibiting very good emulsifying properties. They possess a good phase transfer catalytic ability in Williamson reaction, and their LiClO4 complexes exhibit a conductivity above 1 × 10−4 S/cm at 35°C. © 1996 John Wiley & Sons, Inc.  相似文献   

12.
Series of styrene-methyl methacrylate and styrene-methacrylic acid copolymer latices have been prepared by emulsion polymerisation using polyoxyethylene nonyl phenyl ether as emulsifier and potassium persulphate as initiator. the effects of surfactant concentration and monomer composition on the ultimate particle size and conversion were investigated. The ultimate particle diameters decreased with increasing surfactant concentration, while the conversions were found to be almost independent of surfactant concentration. The ultimate particle diameters were notably decreased by increasing the content of methacrylic acid. Trace carboxyl groups were detected both in polystyrene latex and styrene-methyl methacrylate copolymer latices. The number of sulphate groups on the polystyrene latex surface was about five times that of carboxyl groups.  相似文献   

13.
Photografting reaction onto styrene–butadiene rubber (SBR) as a function of monomer concentration, grafting method, irradiation time, and the carbon black content has been studied using ultraviolet (UV). Glycidyl methacrylate and benzophenone are used as monomer and initiator, respectively. The occurrence of graft reaction onto SBR surface is identified by infrared attenuated total reflection (IR-ATR) analysis. The degree of monomer graft increases with monomer concentration and tends to level off at high monomer concentration (>8.3M/L). Graft ratio also increases with UV irradiation time. Carbon black content is found as one of important factors that determine the monomer graft efficiency. The amount of monomer graft onto SBR decreases with increasing carbon black content and it is attributed to the reduction of irradiation absorbance due to the presence of carbon black. The occurrence of reaction between glycidyl methacrylate grafted SBR and nylon-6 via melt phase reaction is also identified using IR-ATR analysis. © 1999 John Wiley & Sons, Inc. J Appl Polym Sci 73: 1733–1739, 1999  相似文献   

14.
The glass transition temperatures (Tgs) of a series of poly N-(n-alkyl) maleimides covering only the even members with side chains ranging in length from ethyl to n-octadecyl have been studied from room temperature to above Tg. Tgs and thermal quantities have been determined from the specific volume-temperature relations only for the higher (n=8, 10, 12, 14, 16 and 18, where n = no of CH2) members of the series. However for the lower ones (n = 2, 4, 6, 8 and 10) Tgs have been detected from heat capacity-temperature traces of differential scanning calorimetry diagrams by extrapolation to zero rate of heating. Accurate consistency was found in the values (n = 8 and 10) determined by both experimental methods. TgS of these polymers continuously decrease as the number of methylene groups in the side chain is increased, and they have been correlated with the size of the n-alkyl group in the side chain. The results are in accord with a previously studied series concerning the effect of a long side chain on the Tg of a comb-like polymer in the amorphous state. Tgs of poly N-(n-alkyl) maleimides encompassing a wide range of methylene group content (n = 2, 4, 6, 8, 10 and 12) have been examined according to the Gordon-Taylor-Wood extrapolation with the objective of ascertaining the Tg of polyethylene (PE). Our approach of ignoring higher members of the homologous series in this extrapolation appears to be old and well known and it has been variously ascribed to different authors. Extrapolation of Tg values to 100% amorphous PE gives a Tg of 200/pm 10 °K in complete agreement with recent predictions made by Boyer from different sources of data. The Simha-Boyer free volume quantity Δa. Tg decreases slowly with the methylene group content in the longer terms (n = 8, 10, 12, 14, 16 and 18) of the series presumably because of a reduction in the polarity or an in-chain crankshaft loss mechanism which generates free volume in the glassy state, as stated by Boyer. Tgs do not correlate very well with the contributions of the atomic groups to the cohesive energy density (c.e.d.) so it can be concluded that c.e.d. is not the only factor determining Tg. However, a somewhat improved relationship might be obtained by taking into account the steric hindrance effect according to an approach made by Hayes.  相似文献   

15.
16.
The thermal behaviors and the flame‐retardancy of styrene–ethylene–butadiene–styrene–block copolymer containing various additives were studied. The combustion was measured by the Underwriter laboratory (UL) test and cone calorimeter test and thermogravimetric analysis and program‐mass spectroscopy were applied to analyze the thermal behaviors. The blend with halogen additives showed the best result in the UL test. However, the blend with red‐phosphorous was the best in the cone calorimeter test. As the styrene sequence in the copolymer tended to degradate at a lower temperature, the major scission products spouted out from the polymer surface originated from polystyrene. The shorter the ignition times of the blends with red‐phosphorous were, the lower the peak heat release rates were. It was an interesting phenomenon because it suggested that the chemical structure of the residue changed to more stable polymers. © 2007 Wiley Periodicals, Inc. J Appl Polym Sci 104: 156–161, 2007  相似文献   

17.
Acrylonitrile–styrene–methyl methacrylate (A–S–M) terpolymers were prepared by photopolymerization using uranyl nitrate ions as photo initiators, which were analyzed by NMR spectroscopy. The terpolymer compositions were determined by Goldfinger's equation using comonomer reactivity ratios: ras = 0.04; rsa = 0.31; ram = 0.17, rma = 1.45; rsm = 0.52; rms = 0.47. The terpolymer compositions were also determined from the quantitative 13C(1H)‐NMR spectroscopy. The sequence distribution of the acrylonitrile‐, styrene‐, and methyl methacrylate–centered triads were determined from the 13C(1H)‐NMR spectra of the terpolymers and are in good agreement with triad concentrations calculated from the statistical model. © 1999 John Wiley & Sons, Inc. J Appl Polym Sci 74: 3026–3032, 1999  相似文献   

18.
The cationic copolymerization of regular soybean oil, low‐saturation soybean oil (LoSatSoy oil), or conjugated LoSatSoy oil with styrene and divinylbenzene initiated by boron trifluoride diethyl etherate (BF3·OEt2) or related modified initiators provides viable polymers ranging from soft rubbers to hard, tough, or brittle plastics. The gelation time of the reaction varies from 1 × 102 to 2 × 105 s at room temperature. The yields of bulk polymers are essentially quantitative. The amount of crosslinked polymer remaining after Soxhlet extraction ranges from 80 to 92%, depending on the stoichiometry and the type of oil used. Proton nuclear magnetic resonance spectroscopy and Soxhlet extraction data indicate that the structure of the resulting bulk polymer is a crosslinked polymer network interpenetrated with some linear or less‐crosslinked triglyceride oil–styrene–divinylbenzene copolymers, a small amount of low molecular weight free oil, and minor amounts of initiator fragments. The bulk polymers possess glass‐transition temperatures ranging from approximately 0 to 105°C, which are comparable to those of commercially available rubbery materials and conventional plastics. Thermogravimetric analysis (TGA) indicates that these copolymers are thermally stable under 200°C, with temperatures at 10% weight loss in air (T10) ranging from 312 to 434°C, and temperatures at 50% weight loss in air (T50) ranging from 445 to 480°C. Of the various polymeric materials, the conjugated LoSatSoy oil polymers have the highest glass‐transition temperatures (Tg) and thermal stabilities (T10). The preceding properties that suggest that these soybean oil polymers may prove useful where petroleum‐based polymeric materials have found widespread utility. © 2001 John Wiley & Sons, Inc. J Appl Polym Sci 80: 658–670, 2001  相似文献   

19.
Summary Glass transition temperatures of different Polyarylate/plasticizer systems with plasticizer concentrations limited to 30 wt.% have been determined by Differential Scanning Calorimetry. Experimental results have been compared to the predictions of three different expressions: the logarithmic rule of mixtures, the Fox equation and the Couchman-Karasz equation. The degree of homogeneity of the different systems has been related to the width of the simple Tg. The data for the different PAr/plasticizer mixtures seem to fit the Fox equation more closely than the other equations tested.  相似文献   

20.
The impact strength of the acrylonitrile-co-butadiene-co-styrene terpolymer–poly(2,6-dimethyl-1,4-phenylene oxide (ABS–PPO) blends compatibilized with styrene–acrylonitrile modified with glycidil methacrylate (SAN–GMA) terpolymer can be significantly enhanced by the various processing conditions in reaction extrusion. Four different ABS terpolymers are used depending on the composition of acrylonitrile, styrene, and butadiene. The morphology of polybutadiene latex in ABS-1, ABS-3, and ABS-4 is an agglomerated type, while that of ABS-2 is a bimodal one. The three different methods in in situ compatibilizing extrusion are employed; the simple mixing of ABS and PPO, the simultaneous mixing of ABS and PPO, the reactive compatibilizer SAN–GMA, maleic anhydride (MA; designated A-series), and then the stepwise mixing of the mixtures of ABS–SAN–GMA in the MA-modified PPO (designated B-series). Although the ABS-4–PPO blend depicted the highest impact strength in the simple mixing, the ABS-3B–PPO blend showed the best impact strength in the stepwise mixing. The former behavior may be arisen from the high content of BR, whereas the latter may be due to the agglomerated rubber phase with SAN–GMA. The highest impact strength (47 kg·cm·cm−1) was observed in ABS-3B–PPO at 50/50 with an inclusion of 10 wt % GMA (2) and 1 wt % MA. Thus, the proposed reaction mechanism is an existence of the compatibility between ABS and SAN–GMA and the reactivity between the MA-modified PPO and SAN–GMA. Phase morphology of the ABS-2–PPO and ABS-3–PPO blends were compared, and more efficient dispersion of ABS was observed in the B-series than in the A-series. © 1999 John Wiley & Sons, Inc. J Appl Polym Sci 73: 841–852, 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-23

京公网安备 11010802026262号