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 46 毫秒
1.
A series of copolyether macrodiols was prepared from either 1,10-decanediol or 1,6-hexanediol, by acid-catalyzed condensation polymerization using several comonomers to investigate the effect of copolymerization on reducing macrodiol crystallinity. The comonomers used to disrupt crystallinity included 2,2-diethyl-1,3-propanediol, 1,4-cyclohexanedimethanol, and 1,7-heptanediol. The product copolyethers were identified as hydroxy terminated copoly(alkylene oxides) by 1H- and 13C-NMR spectroscopy. Based on NMR results, the structures of the copolyethers were established as consisting of blocks of the principal monomer with comonomer 2,2-diethyl-1,3-propanediol incorporated to form only the end structural unit, whereas 1,4-cyclohexanedimethanol incorporated to form the end unit as well as part of the main chain. DSC results confirmed that the copolymerization produced macrodiols with lower crystallinity and lower Tg than those of the corresponding homopolyethers of the principal monomers, with two exceptions. The exceptions were 1,6-hexanediol/1,10-decanediol, and 1,10-decanediol/1,7-heptanediol copolyethers where no reduction in crystallinity was observed. A series of polyurethane elastomers with a constant hard segment percentage (40 wt %) was prepared using 4,4′-methylenediphenyl diisocyanate and 1,4-butanediol as the hard segment. Tensile test results and Shore hardness measurements demonstrated that copolyether macrodiols produced several polyurethanes with lower modulus and hardness than those of polyurethanes based on homopolyethers of the principal monomers. Of the comonomers studied, 2,2-diethyl-1,3-propanediol-based copolyether produced the polyurethane with the lowest hardness and modulus. © 1997 John Wiley & Sons, Inc. J Appl Polym Sci 63: 1373–1384, 1997  相似文献   

2.
A series of six polyurethanes were prepared to study the effect of silicon chain extender structure on properties and morphology of siloxane–polyurethanes. Polyurethanes were prepared by a two‐step bulk polymerization without a catalyst. The soft segment of the polyurethanes was based on an 80:20 (w/w) mixture of α,ω‐bis(6‐hydroxyethoxypropyl) polydimethylsiloxane (PDMS, MW 966) and poly(hexamethylene) oxide (MW 714). The hard segment was based on 4,4′‐methylenediphenyl diisocyanate (MDI) and a 60:40 molar mixture of 1,4‐butanediol (BDO) and a silicon chain extender. Silicon chain extenders (SCE) investigated were 1,3‐bis(4‐hydroxybutyl)1,1,3,3‐tetramethyldisiloxane (BHTD), 1,3‐bis(3‐hydroxypropyl)1,1,3,3‐tetramethyldisiloxane (BPTD), 1,4‐bis(3‐hydroxypropyl)1,1,3,3‐tetramethyldisilylethylene (HTDE), 1,3‐bis(6‐hydroxyethoxypropyl)1,1,3,3‐tetramethyldisiloxane (BETD). All polyurethanes were clear and transparent with number average molecular weights between 72,000 to 116,000. Incorporation of the silicon chain extender resulted in polyurethanes with low‐modulus and high elongation. This was achieved without significant compromise in ultimate tensile strength in all cases, except BETD. Differential scanning calorimetry (DSC) results showed that the silicon chain extenders did not significantly disrupt the hard segment crystallinity, but exhibited a unique morphological feature where SCE‐based hard segments formed separate domains, which may be the primary reason for achieving low modulus without significant compromise in strength. © 2002 Wiley Periodicals, Inc. J Appl Polym Sci 87: 1092–1100, 2003  相似文献   

3.
1,3-双(3-羟丙基)-1,1,3,3-四甲基二硅氧烷的制备   总被引:3,自引:0,他引:3  
姜红芹  张墩明  蒋锡群  刘芸  杨昌正 《精细化工》2004,21(3):232-234,240
介绍了一种合成1,3 双(3 羟丙基) 1,1,3,3 四甲基二硅氧烷(Ⅰ)的方法,以1,3 双(3 氯丙基) 1,3 二甲氧基 1,3 二甲基二硅氧烷(Ⅱ)为起始原料,其总收率可达75%以上。在25℃将450mL浓度为1mol/L的甲基碘化镁乙醚溶液滴加入70 2gⅡ中,并在40℃回流反应5h,可得到52 6g1,3 双(3 氯丙基) 1,1,3,3 四甲基二硅氧烷(Ⅲ);以N,N 二甲基甲酰胺为溶剂,13 9gⅢ和14 2g醋酸钾在130℃反应7h,生成1,3 双(3 乙酰氧丙基) 1,1,3,3 四甲基二硅氧烷(Ⅳ)15 9g;以碳酸钾为催化剂,在室温下,10 0gⅣ用甲醇醇解1h可制得7 0g标题化合物Ⅰ。用IR、1HNMR和元素分析等分析方法对产物Ⅰ和中间物Ⅲ、Ⅳ等进行了表征。  相似文献   

4.
Acid-catalysed condensation polymerization of 1,6-hexanediol, 1,8-octanediol and 1,10-decanediol was carried out in the 150–190°C temperature range in the presence of sulphuric acid or Nafion-H resin, and the polymerizations were monitored by size-exclusion chromatography. The products were identified as hydroxy-terminated poly(alkylene oxides) by 1H and 13C NMR and IR spectroscopy. The polymers were obtained in 66–81% yield from both catalysts in the polymerizations of 1,8-octanediol and 1,10-decanediol. With 1,6-hexanediol the polymer yields were 30 and 42-56% in the presence of Nafion and sulphuric acid, respectively. These low yields were due to the formation of oxepane as confirmed by gas chromatographic analysis of volatile condensation products. Vapour pressure osmometry showed the number-average molecular weight of polymers to be in the range 700-2400, depending on the reaction time and temperature. The hydroxy functionality of the poly(alkylene oxides) was 2.0 as determined by vapour pressure osmometry and 1H NMR spectroscopy.  相似文献   

5.
Polyester based polyurethanes were synthesized from low molecular weight polyester (Mn2000) and 4,4-methylene bis(phenyl isocyanate) (MDI) with butanediol as a chain extender and tris(6-isocyanatohexyl) isocyanurate (HDT-LV), a tri-NCO terminated compound, as a crosslinker. The polyester was synthesized from adipic acid and glycol, which was a mixture of 1,6-hexanediol and 1,2-propanediol. Two series of crosslinked polyurethanes were prepared. One series were prepared using HDT-LV crosslinked on hard segments and the other series were crosslinked on soft segments. The effect of crosslinker content on the degree of hard segment H-bond formation and phase segregation of polyurethanes was investigated using DSC (differential scanning calorimetry) and FTIR (Fourier transform infrared spectroscopy). The experimental results revealed that the incorporation of tri-NCO crosslinker into the hard-segments of polyurethane resulted in a decrease of hard segment H-bond formation with increasing crosslinker content. However, only a slight decrease of hard segment H-bonding was observed with increasing tri-NCO crosslinker content while it was incorporated into soft segments.  相似文献   

6.
四乙氧基硅烷与烯丙醇在乙醇钠催化下发生酯交换反应,得到一元交换产物三乙氧基烯丙氧基硅烷(Ⅰ)。将Ⅰ与1,1,3,3-四甲基二硅氧烷在铂催化剂条件下进行硅氢加成反应,得到1,3-二(γ-三乙氧硅氧基丙基)-1,1,3,3-四甲基二硅氧烷(Ⅱ),Ⅱ在氢氧化钠水溶液和乙醇以质量比1∶1配成的混合溶液均相水解,得到1,3-双(γ-羟丙基)-1,1,3,3-四甲基二硅氧烷。以1,1,3,3-四甲基硅氧烷计算,总得率为76.4%。采用红外光谱、核磁共振、气相色谱/质谱联用等方法对中间体及产物进行了结构表征。  相似文献   

7.
Three series of polyurethanes were prepared having 42 wt % hard segments based on 4,4′-dicyclohexyl methane diisocyanate (H12MDI) with trans,trans isomer contents in the 13 to 95 mol % range and 1,4-butanediol chain extender. The soft segments were based on macrodiols poly(hexamethylene oxide) (PHMO, MW 696), α,ω-bishydroxyethoxypropyl polydimethylsiloxane (PDMS, MW 940), and two mixed macrodiol compositions consisting of 80 and 20% (w/w) PDMS. H12MDI with 35, 85, and 95% trans,trans isomer contents were obtained from commercial H12MDI (13% trans, trans) by fractional crystallization, and all polyurethanes were prepared by a one-step bulk polymerization procedure. The polyurethanes based on the commercial diisocyanate-produced materials soluble in DMF with molecular weights in the 53,655–75,300 range and generally yielded clear and transparent materials. The polyurethanes based on H12MDI with trans,trans contents of 35% or higher yielded materials insoluble in N,N-dimethylformamide (DMF) and were generally opaque. Mechanical properties, such as tensile strength and elongation at break, decreased with increasing trans,trans content, while the Young's modulus and Shore hardness increased. The polyurethanes based on mixed macrodiols yielded higher tensile properties than those of materials based on individual macrodiols. The best mechanical properties were observed for a polyurethane consisting of a soft segment based on PDMS–PHMO (80/20) and a hard segment based on commercial H12MDI and BDO. Differential scanning calorimetry (DSC) and Fourier transform infrared spectroscopy (FTIR) were employed to characterize the polyurethane morphology. DSC results confirmed that the polyurethanes based on H12MDI with high trans,trans isomer were very highly phase separated, exhibiting characteristic hard segment melting endotherms as high as 255°C. The other materials were generally phase mixed. FTIR spectroscopy results corroborated DSC results. © 1999 John Wiley & Sons, Inc. J Appl Polym Sci 73: 573–582, 1999  相似文献   

8.
The complexes [M(1,3-bis(N-imidazoyl)propyl-1,1,3,3-tetramethyldisiloxane)2(NCS)2] (M = Mn, Ni) are shown by X-ray studies to form monomers containing 16-membered chelate rings, but replacement of the imidazole donor groups by simple amino groups results in the formation of chains of 24-membered rings in [Ni(1,3-bis(3-aminopropyl)-1,1,3,3-tetramethyldisiloxane)2(NCS)2].  相似文献   

9.
Two series of polyurethanes were prepared to investigate the effect of comacrodiol structure on properties and morphology of polyurethanes based on the siloxane macrodiol, α,ω‐bis(6‐hydroxyethoxypropyl) polydimethylsiloxane (PDMS). All polyurethanes contained a 40 wt % hard segment derived from 4,4′‐methylenediphenyl diisocyanate (MDI) and 1,4‐butanediol (BDO), and were prepared by a two‐step, uncatalyzed bulk polymerization. The soft segments were based on an 80/20 mixture of PDMS (MW 967) and a comacrodiol (MW 700), selected from a series of polyethers and polycarbonates. The polyether series included poly(ethylene oxide) (PEO), poly(propylene oxide) (PPO), poly(tetramethylene oxide) (PTMO), poly(hexamethylene oxide), and poly(decamethylene oxide) (PDMO), whereas the polycarbonate series included poly (hexamethylene carbonate) diol (PHCD), poly [bis(4‐hydroxybutyl)‐tetramethyldisiloxy carbonate] diol (PSCD), and poly [hexamethylene‐co‐bis(4‐hydroxybutyl)‐tetramethyldisiloxy carbonate] diol (COPD). Polyurethanes were characterized by size exclusion chromatography, tensile testing, differential scanning calorimetry (DSC), and dynamic mechanical thermal analysis (DMTA). The results clearly demonstrated that the structure of the comacrodiol influenced the properties and morphology of siloxane‐based polyurethanes. All comacrodiols, except PEO, improved the UTS of the polyurethane; PHMO and PTMO were the best polyether comacrodiols, while PSCD was the best polycarbonate comacrodiol. Incorporation of the comacrodiol made polyurethanes more elastomeric with low modulus, but the effect was less significant with polycarbonate comacrodiols. DSC and DMTA results strongly supported that the major morphological change associated with incorporation of a comacrodiol was the significant increase in the interfacial regions, largely through the compatibilization with the hard segment. The extent of compatibilization varied with the comacrodiol structure; hydrophilic polyethers such as PEO were the most compatible, and consequently, had poor mechanical strength. Among the polyethers, PHMO was the best, having an appropriate level of compatibility with the hard segment for substantial improvement in mechanical properties. Siloxy carbonate comacrodiol PSCD was the best among the polycarbonates. © 2000 John Wiley & Sons, Inc. J Appl Polym Sci 78: 1071–1082, 2000  相似文献   

10.
以α,ω-双(γ-羟丙基)聚二甲基硅氧烷(BHPDMS)和聚氧四甲基二醇(PHMO)混合大二醇为软链段;以1,4-丁二醇(BDO)为主要扩链剂,1,3-双(4-羟丁基)-1,1,3,3-四甲基二硅氧烷(BHTD)为次级扩链剂,所有试样中硬链段均由4,4'-二苯基甲烷二异氰酸酯(MDI)和混合扩链剂所构成,且w(硬链段)=40%;采用两步溶液聚合法制备混合大二醇基芳香聚氨酯(PU)弹性体。通过力学性能测试、差示扫描量热法(DSC)和动力学热分析法(DMTA),研究了混合扩链剂中n(BDO)/n(BHTD)比值对该PU弹性体性能的影响。结果表明,当n(BDO):n(BHTD)=3:2时,所得PU弹性体具有优异的综合性能;引入BHTD扩链剂后,破坏了硬链段的有序性。  相似文献   

11.
A series of eight polyurethane elastomers was prepared using a two‐step bulk polymerization procedure to investigate the effect of the siloxane chain extender 1,3‐bis(4‐hydroxybutyl)1,1,3,3‐tetramethyldisiloxane (BHTD) on polyurethane properties and morphology. All polyurethanes were based on 40 wt % hard segment derived from 4,4′‐methylenediphenyl diisocyanate (MDI) and a mixture of 1,4‐butanediol (BDO) and BHTD in varying molar ratios. The soft segment was based on an 80 : 20 (w/w) mixture of the macrodiols α,ω‐bis(6‐hydroxyethoxypropyl)polydimethylsiloxane (PDMS, MW 965) and poly(hexamethylene oxide) (PHMO, MW 714). Polyurethanes were characterized by size‐exclusion chromatography, tensile testing, differential scanning calorimetry, dynamic mechanical thermal analysis, and FTIR spectroscopy. Clear and transparent polymers were produced in all cases with number‐average molecular weights in the range of 90,000 to 111,000. The ultimate tensile strength decreased only slightly (15%), but Young's modulus and flexural modulus decreased by 76 and 72%, respectively, compared with that of the pure BDO extended polyurethanes as the amount of BHTD was increased to 40 mol %. This change resulted in “softer” and more elastic polyurethanes. Polyurethanes with BHTD contents above 40 mol % were more elastic but had poor tensile and tear strengths. A 60 : 40 molar ratio of BDO : BHTD produced a “soft” polyurethane, which combined good tensile strength and flexibility. The DSC and DMTA results confirmed that the incorporation of BHTD as part of the hard segment yielded polyurethanes with improved compatibility between hard and soft segments. IR data indicated that the amount of hard segments soluble in the soft‐segment phase increased with increasing BHTD, contributing to the improved phase mixing. © 2002 Wiley Periodicals, Inc. J Appl Polym Sci 83: 736–746, 2002  相似文献   

12.
Waterborne polyurethane dispersions (PUDs) were synthesized with polycarbonates of 1,6-hexanediol of different molecular weight (500–3000 Da) and their properties, adhesion (Hatch adhesion) and coatings on stainless steel properties (Pencil hardness, Persoz hardness, gloss at 60°, chemical resistance, yellowness index) were characterized. The hatch adhesion of the polyurethane coatings to stainless steel was very good and decreased slightly by increasing the molecular weight of the polycarbonate of 1,6-hexanediol. Both the Pencil and Persoz hardness values of the coatings increased by increasing the hard segments content in the polyurethane, i.e. by decreasing the molecular weight of the polycarbonate of 1,6-hexanediol, whereas the gloss and the yellowness index were lower for the coatings obtained with the polycarbonate of 1,6-hexanediol of molecular weight of 500 Da. Very good chemical resistance against ethanol for all polyurethane coatings on stainless steel plates was obtained but for long time of ethanol in contact with the coating surface the chemical resistance decreased, more markedly for the polyurethane coating obtained with the polycarbonate of 1,6-hexanediol of higher molecular weight. In summary, the segmented structure of the waterborne polyurethane dispersion determined the properties of the polyurethane coatings obtained from them.  相似文献   

13.
Polyester based polyurethanes were synthesized from low molecular weight polyester (Mn 2000) and 4,4′-methylene bis(phenyl isocyanate) (MDI) with butanediol as a chain extender and glycerol as a crosslinker. The polyester was synthesized from adipic acid and glycol which was a mixture of 1,6-hexanediol and 1,2-propanediol. The effect of the crosslinker content on the degree of H-bond formation in the hard segments and the physical properties of polyurethanes were studied by differential scanning calorimetry (DSC), thermal mechanical analysis (TMA), Fourier transform infrared spectroscopy (FTIR) and mechanical testing. The experimental results revealed that incorporation of a triol crosslinker into the hard segments of polyurethane results in a decrease of hard segment H-bond formation. The mechanical data indicate that the mechanical properties of polyurethanes depend on the concentrations of physical and chemical crosslinks and that there is an optimum concentration of triol crosslinker for the tensile stress and elongation properties. © 1998 Society of Chemical Industry  相似文献   

14.
The effect of chain extender structure on properties and morphology of α,ω‐bis(6‐hydroxyethoxypropyl) polydimethylsiloxane (PDMS) and poly(hexamethylene oxide) (PHMO) mixed macrodiol‐based aliphatic polyurethane elastomers was investigated using tensile testing, differential scanning calorimetry (DSC), and dynamic mechanical thermal analysis (DMTA). All polyurethanes were based on 50 wt % of hard segment derived from 4,4′‐methylenecyclohexyl diisocyanate (H12MDI) and a chain extender mixture. 1,4‐Butanediol was the primary chain extender, while one of 1,3‐bis(4‐hydroxybutyl)tetramethyldisiloxane (BHTD), 1,3‐bis(3‐hydroxypropyl)tetramethyldisiloxane (BPTD), hydroquinonebis(2‐hydroxyethyl)ether (HQHE), 1,3‐bis(3‐hydroxypropyl)tetramethyldisilylethylene (HTDE), or 2,2,3,3,4,4‐hexafluoro‐1,5‐pentanediol (HFPD) each was used as a secondary chain extender. Two series of polyurethanes containing 80 : 20 (Series A) and 60 : 40 (Series B) molar ratios of primary and secondary chain extenders were prepared using one‐step bulk polymerization. All polyurethanes were clear and transparent and had number‐average molecular weights between 56,000 and 122,100. Incorporation of the secondary chain extender resulted in polyurethanes with low flexural modulus and high elongation. Good ultimate tensile strength was achieved in most cases. DSC and DMTA analyses showed that the incorporation of a secondary chain extender disrupted the hard segment order in all cases. The highest disruption was observed with HFPD, while the silicon‐based chain extenders gave less disruption, particularly in Series A. Further, the silicon chain extenders improved the compatibility of the PDMS soft segment phase with the hard segment, whereas with HFPD and HQHE, this was not observed. © 1999 John Wiley & Sons, Inc. J Appl Polym Sci 74: 2979–2989, 1999  相似文献   

15.
Treatment of 1-β-(4′-acetylphenyl)vinyl-3-vinyl-1,1,3,3-tetramethyldisiloxane (I) (an AB2 monomer) with dihydridocarbonyltris(triphenylphosphine)ruthenium (Ru) leads to a hyperbranched material, poly[1-β-(4′-acetylphenyl)vinyl-3-vinyl-1,1,3,3-tetramethyldisiloxane] (II). I has been prepared by a Pd catalyzed Heck reaction between 4-bromo-acetophenone and 1,3-divinyl-1,1,3,3-tetramethyldisiloxane. The structure of the soluble hyperbranched material (II) has been determined by 1H, 13C and 29Si NMR, as well as by IR and UV spectroscopy. It has also been characterized by GPC, TGA, DSC and elemental analysis. Polymerization occurs by Ru catalyzed addition of the aromatic C−H bonds which are ortho to the activating acetyl group of I across the C−C double bond of the terminal Si-vinyl group in an anti-Markovnikov manner. Received: 8 September 1997/Revised version: 19 October 1997/Accepted: 20 November 1997  相似文献   

16.
Preparation, Analysis and TLC-Separation of Partial Esters of Fatty Acids with Polybasic Alcohols Direct esterification of 99% capric acid, lauric acid, myristic acid, palmitic acid and stearic acid with ethyleneglycol, diethyleneglycol, thiodiethyleneglycol, triethyleneglycol, 1,2-propanediol, 1,2-butanediol, 1,3-butanediol, 1,4-butanediol, 1,5-pentanediol, 2,5-hexanediol, 1,6-hexanediol, 1,8-octanediol, 1,10-decanediol, 1,12-dodecanediol, 1,2,4-butanetriol, glycerol, 1,2,6-hexanetriol, trimethylolpropane and pentaerythritol in a molar ratio of 1 :1.25 yields 100 different partial esters of fatty acids. These partial esters are extensively freed from the polyhydric alcohols by washing with sodium sulfate solution and recrystallization in ethanol. Chemical constants and TLC-separation into classes and into polyhydric alcohols permit the evaluation of these compounds as emulsifiers, stabilizers and solubilization acids. All the partial esters of the fatty acids are mixtures, which can be separated by TLC according to the polyhydric alcohol moiety into monoester, diester, triester and tetraester; furthermore, the separation of positional isomers of monoesters and diesters is possible. Several observations made during the synthesis and analysis of partial esters of fatty acids are reported.  相似文献   

17.
Silicone polyimides were made with siloxane-equilibrated 1,3-bis (3,4-dicarboxyphenyl) - 1,1,3,3-tetramethyldisiloxane dianhydride (PADS) as the soft portion and [(1-methylethylidene) bis (4,1-phenyleneoxy)] bis-1,3-isobenzofurandione (BPADA) as the hard portion in combination with either para or meta-phenylene diamine. The silicone polyimides were made with varying siloxane chain lengths and with varying overall weight percent silicone. There were linear correlations between these variables and flexural modulus, percent elongation at break, and cut-through temperature. Glass transition temperatures were evaluated by dynamic mechanical analysis and the thermal stabilities were determined to be greater than 400°C by thermal gravimetric analysis. Melt viscosities were determined on the various copolymers and demonstrated a correlation between type of amine used, siloxane chain length, and overall weight percent silicone.  相似文献   

18.
以甲苯二异氰酸酯(TDI)、1,4-丁二醇(BDO)为硬段,聚碳酸酯二醇(PCDL)和聚醚二醇(PPG)混合物为软段,采用预聚体法制备了不同软段组成的聚氨酯弹性体(PUE)。采用DSC、FT—IR和DMA等分析手段研究了PCDL含量对PUE的微相分离程度和阻尼性能的影响。结果表明,随着软段中PCDL含量的增加,PUE中氨酯羰基的氢键化程度减小,相分离程度减小,而且PUE的储能模量随着PCDL含量的增加而减小;与单一组分软段的PUE相比较,混合软段的PUE具有相对较好的阻尼性能。  相似文献   

19.
A series of novel silicone-modified polyesters (SPE) were prepared by substituting part of diol with low molecular weight hydroxyl-terminated poly(dimenthylsiloxane) (PDMS). Then, isophorone diisoclanate (IPDI) as hard segments and 1,4-butanediol as chain extender were added to SPE to prepare a silicone-modified polyurethane (SPU). The effects of the type of diol, diacid, and hydroxyl-terminated PDMS, and the amount of hydroxyl-terminated PDMS on the preparation and surface properties of SPU were investigated. It was found that the amount of PDMS incorporated into a polyester chain was relatively higher when 1,6-hexanediol (HDO) and 1,10-decandiol (DDO) were used as diol and the PDMS with lower molecular weight was used as organosilicone compound. Consequently, the SPU coats with HDO as diol, adipic acid (AA) as diacid, and short chain PDMS as silicone segment had the lowest surface-free energy since it had the highest and most homogeneous distribution of silicone segments at its top layer surfaces.  相似文献   

20.
Three waterborne polyurethane dispersions derived from polyester, polyether and polycarbonate diols with molecular weight of 1000 Da were synthesized by the acetone method and used as coatings on stainless steel 304 plates. The properties of the dispersions and the polyurethane films were influenced by the polyol nature. The polyurethanes obtained with polyether or polyester showed higher degree of phase separation between the soft and the hard segment. The higher adhesive strength under shear stresses was obtained in the joints produced with the waterborne dispersion obtained with polycarbonate diol. The properties of the polyurethane coating obtained with polycarbonate diol on stainless steel 304 were significantly higher as compared with the others. Improved performance of coatings obtained with polycarbonate diol was ascribed to the higher polarity of the carbonate groups that contributed to additional hydrogen bond formation between soft segments with respect to those obtained with polyether or polyester  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