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.
A range of reactor grade TPO alloys (PP/EPM alloys from Ziegler‐Natta catalysis) were studied in order to evaluate their dynamic cross‐linkability, using a peroxide system. The amount of ethylene content in EPM rubber ranged from 28 to 77 wt%, to investigate the competition between vis‐breaking and cross‐linking during the peroxide reaction. Processing temperature and different peroxide/coagent ratios were the parameters varied in the present work. Cross‐linking degree was evaluated with Compression Set measurements and Dynamic Mechanical Analysis. It was therefore possible to find a correlation between the two analytical techniques. Feasibility of TPV rubbers with Compression Set at about 50% (22 h, 70°C) was demonstrated.  相似文献   

2.
The scope of this article is the study of peroxide curing of two nitrile rubbers with low and high nitrile content. The peroxide efficiency can be much higher than one, and the polymer structure determines the mechanism of cross‐linking. In the rubber with low nitrile content, the peroxide radical may give rise to a polymerization reaction between adjacent double bonds generating a heterogeneous network with a negative effect on the vulcanizate properties. On the contrary, in the nitrile rubber with high nitrile content, this negative effect it is not present or is present to a lesser extent, and their vulcanizates show good physical properties. © 2005 Wiley Periodicals, Inc. J Appl Polym Sci 96: 1–5, 2005  相似文献   

3.
Sulfur can be used as crosslink coagent in unsaturated elastomer. In this work, a fully saturated HNBR with 39 wt % nitrile content was selected to investigate the effect of a small amount of sulfur acting as crosslink coagent on the physical and mechanical properties of peroxide‐cured vulcanizates. First, selective cleavage of polysulfide (? Sx? ) and monosulfide(? S? ) bond by combined thiol‐piperidine treatment were performed and the existence of poly/monosulfide bond in sulfur‐contained HNBR compounds was verified. Then, no‐filler HNBR compounds with various content of sulfur were investigated to detect the influence of sulfur on the crosslink density and cure kinetics. The MDR results showed that the crosslink density of HNBR compounds reduced only when the amount of sulfur is 0.25 phr and above. Besides, the curing rate of no‐filler HNBR compounds increased with the increasing of the amount of sulfur and reached a maximum at a dosage of 0.25 phr sulfur. Finally, physical and mechanical properties of fully formulated compounds were evaluated and it was found that the addition of small amounts of sulfur in fully saturated HNBR compounds could improve the dynamic properties of peroxide‐cured HNBR compounds remarkably but at a cost of slightly higher compression set values and a small loss in heat aging resistance. In a conclusion, small amount of sulfur can impart the peroxide vulcanizates some “sulfide properties” like dynamic property, tensile strength, but at the same time, due to the introduction of sulfur, some “peroxide vulcanizates property” like heat‐resistance property, hot air resistance were weakened slightly. © 2014 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2015 , 132, 41612.  相似文献   

4.
Novel films based on cross‐linked hyperbranched poly(amine‐ester) (HPAE) were prepared by cross‐linking the terminal hydroxyl groups of HPAE using glutaraldehyde (GA). The cross‐linking process was monitored by measuring the intrinsic viscosity of HPAE/GA in N,N‐dimethylacetamide. The surface structures of the cross‐linked HPAE films obtained from different HPAE/GA ratios were imaged using atom force microscopy, and their properties were characterized in terms of hydrophilicity, solvent swelling, mechanics, and protein adsorption. It was found that the static contact angle was <32.9°, tensile strength was >0.35 MPa, elongation at break was >9.2%, swelling degree was >63% in water, and bovine serum albumin adsorption was relatively low. The results indicate that cross‐linked HPAE films have a strong application potential in many areas. Copyright © 2005 Society of Chemical Industry  相似文献   

5.
High‐density polyethylene was functionalized using a thermolysis method in the presence of functionalized peroxides at different temperatures and at various peroxide concentrations. It was found that both percentage cross‐linking (% CL) and percentage functionalization (% Fn) increased with an increase in peroxide concentration. The ester and epoxide functionalization was confirmed by FTIR spectroscopy. Ester functionalization was further confirmed by saponification and acidolysis reaction. The functionalization capacities of acrylic ester peroxide and acrylic acid peroxide were determined and compared. A plausible reaction mechanism has been proposed to explain the experimental results obtained. © 2005 Wiley Periodicals, Inc. J Appl Polym Sci 97: 761–765, 2005  相似文献   

6.
Poly(3‐hydroxybutyrate) (PHB) and poly(l ‐lactic acid) (PLLA) were individually cross‐linked with dicumyl peroxide (DCP) (0.25–1 wt %) by reactive melt processing. The cross‐linked structures of the polymer gel were investigated by nuclear magnetic resonance (NMR) and Fourier transform infrared (FTIR) spectroscopies. The size of the polymer crystal spherulites, glass transition temperature (Tg), melting transition temperature (Tm), and crystallinity were all decreased as a result of cross‐linking. Cross‐linking density (νe) was shown to increase with DCP concentration. Based on parallel plate rheological study (dynamic and steady shear), elastic and viscous modulus (G″ and G′), complex viscosity (η*) and steady shear viscosity (η) were all shown to increase with cross‐linking. Cross‐linked PHB and PLLA showed broader molar mass distribution and formation of long chain branching (LCB) as estimated by RheoMWD. Improvements in melt strength offer bioplastic processors improved material properties and processing options, such as foaming and thermoforming, for new applications. © 2014 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2015 , 132, 41724.  相似文献   

7.
The influence of cross‐linking on fracture properties of polyethylene has been studied by the method of essential work of fracture (EWF). Three distinct values of the cross‐link density were obtained by β irradiation with three distinct radiation doses, and characterized by sol‐gel and solvent swelling measurements. EWF tests were performed at speeds of 0.045 to 100 mm min–1 at 80 and 110°C. The fracture toughness, as defined from the EWF method, is analyzed through essential (we) and nonessential (βwp) components. Both values were found to be decreasing functions of the cross‐link density. But cross‐linking also suppresses the toughness decrease observed in linear PE at low speeds/high temperatures. It is hypothesized that cross‐linking prevents the chain disentanglement that occurs during crack growth. In conclusion, EWF tests performed at low speeds appear as an interesting method to characterize the influence of structural factors on the fracture properties of polyethylene. POLYM. ENG. SCI., 45:424–431, 2005. © 2005 Society of Plastics Engineers  相似文献   

8.
A series of polymer networks of varying cross‐linking density was prepared by copolymerization of methyl acrylate and ethyleneglycol dimethacrylate. The aim of this work is to study the influence of cross‐linking on the conformational mobility of the polymer chains using dielectric relaxation spectroscopy (DRS) in the temperature range of the main dielectric relaxation. As expected, the temperature range in which glass transition takes place became wider with increasing crosslinking density. DRS results were analyzed using the Havriliak‐Negami equation. Master Cole‐Cole arcs could be drawn for all the networks. The arcs become more symmetric as cross‐linking density increases, as a consequence of the different effect of cross‐links on large and small scale mobility. The conformational mobility that produces the main relaxation is drastically reduced when the cross‐linking density increases what reduces the relaxation strength, but it also gives a qualitative change of behavior, as shown by the temperature dependence of the relaxation strength. In the loosely cross‐linked networks the relaxation strength decreases monotonously as temperature increases, as in the main dielectric relaxation of linear polymers. Nevertheless, in highly cross‐linked networks the curve of relaxation strength against temperature presents a maximum. POLYM. ENG. SCI., 45:1336–1342, 2005. © 2005 Society of Plastics Engineers  相似文献   

9.
Ethylene‐propylene‐diene‐monomer (EPDM) rubber is an important commercial polymer. The vulcanization process significantly changes its thermal, mechanical, and vulcametric properties. This study was carried out to find optimum formulation of EPDM composite for a better application in automotive industry. Sixteen EPDM polymer samples having different 5‐ethylidene 2‐norbornene (ENB) and ethylene contents were vulcanized with different types and compositions of peroxide and coagents. The mechanical and vulcametric properties of these samples were measured and compared. The type of peroxide, coagent, and EPDM grade affected the mechanical, thermal, and vulcametric properties of the EPDM rubber to some extend. Use of aromatic peroxide and coagent increased the thermal stability slightly. Mechanical properties were changed very slightly with the change of peroxide type for the same content of peroxide and coagent. Scorch time and cure time decreased with initial increase of the peroxide content. EPDM compound vulcanized with BBPIB peroxide and TAC/S coagent has higher cure time than EPDM compound vulcanized with DMBPHa peroxide and TMPTMA coagent. Coran method was used for the modeling of experimental data. Velocity constant for the formation of peroxide radical and polymer radical were found for each case. © 2009 Wiley Periodicals, Inc. J Appl Polym Sci, 2009  相似文献   

10.
High‐density polyethylene was thermolysed with short‐chain organic compounds (additives) with an ester functional group in the presence of free radical initiator such as dicumyl peroxide at 160°C. The experimental results inferred that an additive with a hydrogen donor containing a thiol group showed better efficiency toward functionalization and an additive with a methylene bridged group showed better efficiency toward cross‐linking. The FTIR spectral results indicated that the area of the carbonyl peak at 1730 cm?1 increased with the increase in percentage grafting of additives with simultaneous increase of cross‐linking. The ester grafting was further confirmed by saponification reaction. The possibility of acid functionalization via intramolecular hydrogen transfer was also checked through acid hydrolysis of the saponified products. A suitable reaction mechanism was proposed in order to explain the experimental and spectral data obtained. © 2005 Wiley Periodicals, Inc. J Appl Polym Sci 97: 766–774, 2005  相似文献   

11.
Peroxide curing of brominated butyl rubber (BIIR) is an attractive topic, but the degradation of BIIR during the curing is a drawback needed to be overcome. Coagent assisted peroxide curing system is an attractive and effective choice in order to increase the crosslink density of rubbers. 1,2‐polybutadiene (1,2‐PB) is used as a crosslinking coagent for the curing of BIIR by dicumyl peroxide (DCP), and the effect of 1,2‐PB on the curing characteristics, crosslink density, and mechanical properties is investigated. The addition of 1,2‐PB affects the curing characteristics of BIIR compound and significantly increases the crosslink density of BIIR vulcanizates. With increasing 1,2‐PB content, the tensile strength and stresses at a given extension of BIIR vulcanizates increase, but the elongation at break decreases. A stress‐softening effect of the carbon black filled BIIR vulcanizates is observed and becomes more pronounced with increasing 1,2‐PB content. The addition of 1,2‐PB increases the stress relaxation index of BIIR. GPC and 13C‐NMR results indicate 1,2‐PB participates in the crosslinking reaction, and the existence of 1,2‐PB component in the insoluble fraction of BIIR/1,2‐PB vulcanizates is confirmed by solid‐state 13C‐NMR. © 2015 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2016 , 133, 43280.  相似文献   

12.
The cross‐linking reaction of polyethylene/peroxide system was studied at elevated temperature. The dynamic modulus evolution was monitored within linear viscoelastic regime by parallel plate rheometer. Different strains were chosen to represent different flow fields. Based on the mechanism of cross‐linking reaction, a new expression relating recombination rate constant with rheological conversion was derived. The experimental and calculated results showed that the strain had substantial effects on the cross‐linking reaction kinetics. Consequently a formula reflecting the effect of the complex viscosity on recombination rate constant was found, which was further integrated into the classical Arrhenius kinetic equation. POLYM. ENG. SCI., 45:560–567, 2005. © 2005 Society of Plastics Engineers  相似文献   

13.
Triazole cross‐linked polymers based on poly(3‐azidomethyl‐3‐methyl oxetane) (poly‐AMMO) and glycidyl azide polymer (GAP) were prepared using bis‐propargyl‐1,4‐cyclohexyl‐dicarboxylate (BPHA) as curing agent, respectively. Swelling tests demonstrated that cross‐linking densities of the resulted polymers both increased with the increase of BPHA. Triazole cross‐linked polymers based on poly‐AMMO showed superior tensile strength and elongation at break than those of GAP at comparable stoichiometry. The curing kinetics was also investigated by FTIR, and GAP exhibited faster reaction rate when reacted with BPHA than that of poly‐AMMO. In addition, with the increase of cross‐linking density, the glass transition temperature (Tg) of as‐prepared polymers significantly increased, and poly‐AMMO‐based polymers showed stronger Tg‐raising effect than GAP‐based polymers. © 2016 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2016 , 133, 43341.  相似文献   

14.
Developing green and sustainable alternative materials to replace petroleum based ones is the need of the day. Such green materials are becoming popular because they can be composted once their useful life is over. In the current research, protein‐based nanofibers were fabricated without the use of any toxic cross‐linking agent. Defatted soy flour was purified using an acid‐wash process to obtain material with higher protein content, blended with gluten, and successfully electrospun into nanofibers with the help of polyvinyl alcohol. Oxidation of sucrose with hydrogen peroxide (H2O2) was carried out to synthesize oxidized sugar‐containing aldehyde (—CHO) groups and used as green cross‐linker. The cross‐linking quality of protein‐based nanofibers modified by oxidized sugar was found to be similar to nanofibers cross‐linked using toxic glyoxal and show good resistance to water. These novel green protein‐based nanofibers can be useful in fabricating inexpensive products with very high specific surface area and highly porous structure. © 2015 The Authors Journal of Applied Polymer Science Published by Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2015 , 132, 41852.  相似文献   

15.
In this article, mechanical and compression set properties, swelling property, and stress–strain behavior during compression of polysulfide sealants based on different polysulfide resin were investigated. The results showed that molecular weight and cross‐linking agent of liquid polysulfide resin had significant influence on mechanical and compression set properties of the sealants. The sealants based on higher molecular weight polysulfide resin had higher mechanical properties. At the same time, lower cross‐linking agent in polysulfide resin produced lower cross‐link density and higher swelling property, which resulted in higher compression set value of the sealant. However, when different molecular weight polysulfide resins were used in the sealant simultaneously, the testing results indicated that the compression performance of the sealants was significantly enhanced, while mechanical properties of the sealants kept nearly unchanged. © 2009 Wiley Periodicals, Inc. J Appl Polym Sci, 2010  相似文献   

16.
The objective of this study is to investigate the effects of carbon nanotube (CNT) content, surface modification, and silane cross‐linking on mechanical and electrical properties of linear low‐density polyethylene/multiwall CNT nanocomposites. CNTs were functionalized by vinyltriethoxysilane to incorporate the ─O─C2H5 functional groups and were melt‐blended with polyethylene. Silane‐grafted polyethylene was then moisture cross‐linked. Silanization of CNT was confirmed by Fourier transform infrared spectroscopy, thermogravimetric analysis (TGA), and EDX analysis. Hot‐set test results showed that silane cross‐linking of polyethylene and incorporation of modified CNTs into polyethylene led to an increase in cross‐linking density and the number of entanglements resulting in a decrease in elongation. It was found that the addition of pristine multiwall carbon nanotubes (MWCNTs) and functionalized MWCNTs does not affect silane cross‐linking density. Silane modification resulted in a stronger adhesion of the silane cross‐linked LLDPE to silanized MWCNTs according to scanning electron microscopy micrographs. Additionally, the electrical tests revealed that the silane modification of CNTs results in an improvement in electrical properties of nanocomposites, while silane cross‐linking will not have an effect on electrical properties. Rheological properties of MWCNT/LLDPE nanocomposites have been studied thoroughly and have been discussed in this study. Moreover, according to TGA test results, modification of the MWCNTs led to a better dispersion of them in the LLDPE matrix and consequently resulted in an improvement in thermal properties of the nanocomposites. Crystallinity and melting properties of the nanocomposites have been evaluated in detail using DSC analysis. J. VINYL ADDIT. TECHNOL., 26:113–126, 2020. © 2019 Society of Plastics Engineers  相似文献   

17.
This study aims at obtaining cross‐linked polymeric materials of biomass origin. For this purpose, one‐pot polyesterification of methyl ricinoleate and methyl 12‐hydroxystearate using titanium isopropoxide as a catalyst is performed leading to polyesters known as estolides. The obtained estolides are successfully cross‐linked using dicumyl peroxide or a sulfur vulcanization system. The so‐formed bio‐based elastomers appear to exhibit promising properties. The latter are analyzed by mechanical tensile tests and thermal techniques (TGA, DSC, DMA) and show high thermal stability (T5% = 205–318 °C), tailored physico‐mechanical properties (low glass transition temperature in the range from ?69 to ?54 °C), and good tensile strength (0.11–0.40 MPa). Networks prepared from high molecular weight estolides appear to be promising bio‐based elastomers. Practical Applications: The vegetable oil‐based estolides described in this contribution are new fully bio‐based precursors for further elastomers synthesis. The resulting estolide networks (obtained by peroxide or sulfur cross‐linking) exhibit tailored thermo‐mechanical properties.  相似文献   

18.
Two series of hydrogels of poly[(acrylic acid)‐co‐(itaconic acid)] have been prepared by copolymerization in solution using tetrafunctional N,N′‐methylenebisacrylamide (NMBA) as cross‐linker. The resulting polymer was swollen in water at 298 K to yield homogenous transparent hydrogels. These hydrogels were characterized in terms of swelling and compression‐strain measurements. The influence of the comonomer composition and concentration of cross‐linking agent on volumetric swelling and the mechanical properties of these hydrogels were investigated. Inefficient cross‐linking is indicated by the small values of νe relative to the theoretical cross‐linking densities.

Dependence of measured affective cross‐linking density (νe) on the theoretical cross‐linking density (νt) for acrylic acid/itaconic acid hydrogels prepared at a fixed composition of AA80/AI20 wt.‐%, but at different concentrations of NMBA.  相似文献   


19.
Blend of ethylene‐vinyl acetate copolymer (EVA), ethylene‐propylene‐diene terpolymer rubber (EPDM), and chlorinated polyethylene (CPE) filled with talc was cross‐linked with dicumyl peroxide and blown simultaneously with azodicarbonamide for preparation of high elasticity material by injecting, cross‐linking, and blowing process. It had excellent mechanical properties and elasticity. It was found that mechanical properties were a function of cross‐linking and blowing temperature and time and formula. A good cross‐linking and blowing temperature was in the range of 440–450 K and the optimum temperature and time were 448 K and 360 s, respectively. Kinetics of both cross‐linking and blowing reactions were investigated by closed nonrotor torque rheometer. The results showed that both cross‐linking and blowing reactions were first order. The activation energy of the cross‐linking reaction was higher than that of the blowing reaction. With increasing of EDPM content, the activation energy of the cross‐linking reaction reduced whereas it increased for the blowing reaction. The gas cell morphology was characterized by scanning electron microscopy (SEM). The results of SEM and mechanical property analyses showed that the morphology and mechanical properties were improved by incorporating CPE. © 2007 Wiley Periodicals, Inc. J Appl Polym Sci, 2007  相似文献   

20.
A series of polyurethane (PU) polymers cross‐linked laterally by pentaerythritol and polyethyleneglycol (PEG) spacers were compared with linear PU. The PU was composed of 4,4′‐methylenebis(phenylisocyanate) (MDI), poly(tetramethyleneglycol), 1,4‐butanediol (BD), pentaerythritol, and PEG‐200 as a spacer. PEG‐200 connected the pentaerythritol hydroxyl groups of two PU chains with MDI as a connecting agent. The phase separation between hard and soft segments was disrupted by the PEG crosslinking, and Tm did not change with an increase in cross‐linking content. Instead, the cross‐link density increased with an increase of pentaerythritol content. A significant increase in maximum stress compared with linear PU was attained, together with an increase in strain. The combination of both pentaerythritol and PEG‐200 in the PU resulted in the improvement of both stress and strain, unlike in the conventional cross‐linking method. The shape recovery increased to 90% and did not decrease after three test cycles. © 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