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 451 毫秒
1.
《塑料》2015,(3)
采用两种酸一种醇和两种醇一种酸分别对PBS改性,合成了不同化学结构的共聚酯PBSA和PBSH,并在磷酸缓冲液中以它们为底物,研究了对脂肪酶N435降解反应的异同。采用质量损失率和GPC评价了降解前后共聚物分子质量的变化;WAXD和TG分析了酶降解前后共聚物结晶度和热性质的变化;POM对降解后的材料进行了形貌观测。研究结果表明:相比PBS,PBSA和PBSH对脂肪酶的感受性有很大提高,24 h后降解率分别达到90%和60%以上,并且PBSA降解速率比PBSH快很多;降解后2种共聚物相对分子质量变化不大,但分子量分布系数变宽;结晶度增大;降解3 d后PBSA的热稳定性降低,而PBSH的热稳定性提高。  相似文献   

2.
采用L_9(3~3)正交试验对α–淀粉酶降解聚丁二酸丁二酯(PBS)基共聚酯/热塑性淀粉(TPS)复合材料的条件进行了优化,得出α–淀粉酶的最优降解条件为:温度65℃,磷酸盐缓冲液p H=6.8,α–淀粉酶浓度3.5 g/L。利用α–淀粉酶和南极假丝酵母脂肪酶N435对PBS/TPS、聚(丁二酸丁二醇-co-丁二酸二甘醇)酯(PBS-co-DEG)/TPS、聚(丁二酸丁二醇-co-丁二酸乙二醇-co-丁二酸聚乙二醇200)酯(PBES-co-PEG200)/TPS、聚(丁二酸丁二醇-co-丁二酸乙二醇-co-丁二酸聚乙二醇400)酯(PBES-co-PEG400)/TPS复合材料分别进行降解实验,研究了两种酶对这4种复合材料降解性能的影响。结果表明,α–淀粉酶和N435脂肪酶对复合材料均有较好的降解能力,当降解时间较短(6 h)时,α–淀粉酶对复合材料的降解效果优于脂肪酶N435,但当降解时间超过60 h后,后者的降解效果略优于前者;(PBES-co-PEG200)/TPS和(PBES-co-PEG400)/TPS复合材料的降解性能总体上优于(PBS-coDEG)/TPS及PBS/TPS复合材料;随PEG200和PEG400在共聚酯中的含量增加,即醚链含量的增加,相应复合材料的质量损失率呈升高趋势,但当醚链含量较高时,复合材料的质量损失率反而有所下降。  相似文献   

3.
聚丁二酸丁二醇酯(PBS)是一种具有广阔发展前景的生物基聚酯品种,但由于其分子链段柔性较高并且结晶度高,在实际应用过程中仍然面临热性能、力学性能、降解性能不足的弊端。选用条件更为温和的酶催化体系,以丁二酸二乙酯、己二酸二乙酯、丁二醇为原料,Novozym-435(固定化CALB酶)为催化剂,甲苯为溶剂,合成制备了一系列高分子质量的聚(丁二酸丁二醇-co-己二酸丁二醇)(PBSA)三元共聚酯。在反应温度80℃、酶的质量分数为10%的条件下,所制备的PBS与PBSA共聚酯的数均分子质量在18 400~21 200 g/mol,多分散性指数在1.7~2.0之间。采用核磁共振谱仪对制备的共聚酯产物的化学结构进行了表征。TGA结果表明,酶催化体系下的PBSA共聚酯的热稳定性要高于熔融体系下的产物。DSC与WAXD的结果表明,引入己二酸链段后,PBS的结晶能力下降。酶催化反应避免了重金属催化剂的使用,进一步拓宽了PBS在生物医药领域范围的潜在应用。  相似文献   

4.
韩艳萍  杨景辉 《塑料工业》2012,40(12):29-32,45
采用直接熔融缩聚法,用丁二酸和丁二醇分别与1,2-丙二醇、1,2-戊二醇、1,2-己二醇共聚改性合成得到一系列产物聚丁二酸丁二醇酯(PBS)、聚(丁二酸丁二醇酯-co-丁二酸1,2-丙二醇酯)(PBSP)、聚(丁二酸丁二醇酯-co-丁二酸1,2-戊二醇酯)(PBST)和聚(丁二酸丁二醇酯-co-丁二酸1,2-己二醇酯)(PBSH).利用乌氏黏度计、1H NMR、DSC等对其摩尔质量、化学结构、热学性能和力学性能进行表征.结果表明,随着共聚酯分子主链上支链长度的增加,数均摩尔质量(Mn)几乎无变化,对应的熔点(Tm)、结晶温度(Te)、结晶度(Xc)、弯曲强度和拉伸强度逐渐降低,断裂伸长率明显增加.冲击强度变化:PBSP-10< PBST-10< PBSH-10<PBS,总体上PBSH-10表现出良好的综合力学性能.  相似文献   

5.
马莹  侯微  苏婷婷 《塑料科技》2023,(1):95-100
通过酯化和缩聚反应制备聚丁二酸丁二醇酯(PBS)、聚己二酸丁二醇酯(PBA)和聚(丁二酸-co-己二酸丁二醇)共聚酯(P(BS-co-BA)),对PBS、PBA和P(BS-co-BA)进行酶促降解研究。结果表明:与PBS和PBA相比,共聚酯具有良好的生物降解性能。6种聚酯酶水解速率依次为P(BS-co-40%BA)>P(BS-co-60%BA)>P(BS-co-80%BA)>P(BS-co-20%BA)>PBA>PBS。P(BS-co-40%BA)在10 h内基本完全降解,比PBS快26 h。与PBA相比,共聚酯的热稳定性得到提高,P(BS-co-40%BA)热分解50%的温度比PBA高22.3℃。随着降解时间的增加,共聚酯的化学结构、晶体结构和热稳定性基本不变,有利于其在新能源汽车设计中的应用。  相似文献   

6.
介绍了聚丁二酸丁二醇酯(PBS)基聚酯的生物降解研究及相关影响因素,分别从微生物降解,生理环境降解,酶降解三方面进行总结,并对聚丁二酸丁二醇酯(PBS)基聚酯降解的研究方向及应用前景进行了展望。  相似文献   

7.
徐秋舒  陈文生  崔燕军 《塑料》2024,(1):109-113
以1,4-丁二酸(SA)、1,4-丁二醇(BDO)、1,2-环己二醇(CHD)为原料,通过改变BDO与CHD投料比,采用熔融缩聚法制备了一系列的CHD改性聚丁二酸丁二醇酯(PBS)共聚酯。采用1H-NMR表征了共聚酯的化学结构,并且,分析了CHD占主链二醇含量对共聚物分子量及其分子量分布、熔融和结晶性能、热稳定性能、拉伸性能及脂肪酶降解性能的影响。结果表明,随着CHD含量的增加,共聚酯的数均分子量从7.45×104下降至4.75×104,由结晶度48.5%的半晶态转变为无定形态,热分解损失为质量5%时,温度降低了26.7℃,拉伸强度由38.2 MPa降低至14.9 MPa,但脂肪酶降解性能显著提高。PBS主链引入适量CHD后,可以有效地调控PBS的结晶度及柔顺性,提高了PBS在非堆肥条件下的降解速度。  相似文献   

8.
张敏  覃家祥  李成涛  张祎  邱建辉 《塑料》2014,(4):91-94,113
采用己二醇对聚丁二酸丁二醇酯(PBS)改性,合成了不同比例、分子质量均在6×105左右的聚丁二酸丁二醇/己二醇酯(PBSH),并以其为底物分别在2种不同的有机溶剂氯仿(CHCl3)和四氢呋喃(THF)中,研究了洋葱假单胞菌(PC)脂肪酶对其催化降解规律和溶剂效应。以GPC测试了共聚物降解前后的分子质量变化;以TG分析了酶降解共聚物前后热性能的变化;以MALDI-TOF-MS对降解产物进行了分析。研究结果表明:PBSH在2种溶剂中都能快速降解;降解60 h后2种共聚物的相对分子质量均减小,分子质量分布均变宽;但在氯仿中酶催化活性更高,PBSH降解速率更快;降解前后热失重5%时热分解温度均降低;MALDI-TOF-MS结果表明:在2种溶剂中降解产物中含SH(丁二酸己二醇酯)片段较多,且氯仿中降解产物种类更多,并易于成环。  相似文献   

9.
《塑料》2019,(6)
为降低PBS的降解速率,以不同聚合度的聚碳酸酯二醇(PCDL)对聚丁二酸丁二醇酯(PBS)进行了共聚改性。1H-NMR和GPC结果表明,通过熔融缩聚成功合成了数均分子量Mn大于46 000的聚(丁二醇丁二酸-co-聚碳酸酯二醇)共聚酯P(BS-co-CDL)。DSC和TG测试结果表明,PCDL的加入使材料的结晶度、熔点及结晶温度降低,耐热性提高。力学性能测试结果表明,PCDL的加入使材料的强度和模量降低,断裂伸长明显增加。其中,4#2 000-3%样品的拉伸强度为38. 85 MPa,弹性模量为263. 13 MPa,断裂伸长率为455. 84%。65℃磷酸盐缓冲液中降解性能测试表明,与PBS相比,PCDL的引入使P(BS-co-CDL)共聚酯的降解速率较慢,其中,4#2 000-3%样品的降解速率最慢,36 d后,失重率为3. 21%,此时,P(BS-co-CDL)的综合性能最佳。  相似文献   

10.
采用多种二醇改性聚丁二酸丁二醇酯(PBS),合成了碳链长度不同的PBS基共聚酯。在CHCl3中,以固定化南极假丝酵母脂肪酶b(N435)降解各共聚酯,研究了碳链长度对共聚酯降解速率和降解率的影响。通过共聚酯分子量、降解产物及热稳定性变化,分析了影响降解效果的因素。采用分子对接技术解释了脂肪酶与底物之间的相互作用机理。结果表明:随共聚酯碳链增长,共聚酯降解效果逐步提升,己二醇改性时的降解速率最高,降解率可达80%;共聚酯分子链越长,产生的低聚物越多;随分子链增长,共聚酯热稳定性逐步下降。分子对接显示:脂肪酶催化三联体、活性口袋残基及底物三者之间形成的氢键对酶的催化作用至关重要。  相似文献   

11.
采用长碳链癸二酸代替丁二酸与丁二醇以及不同含量的相对分子质量为200的聚乙二醇(PEG)嵌段共聚,得到了聚癸二酸丁二醇酯聚乙二醇嵌段共聚物(PBSe-PEG),并采用核磁共振波谱仪、凝胶渗透色谱仪、差示扫描量分析仪、热重分析仪、广角X射线衍射仪等分析手段表征了嵌段共聚物的结构和性能,并且研究了固定化南极假丝酵母菌脂肪酶(简称脂肪酶N435)对PBSe-PEG的降解性。结果表明,几种PBSe-PEG是预期产物,其数均相对分子质量在5×104左右;PBSe-PEG与纯PBSe相比晶型结构相似;随着PEG含量增加,PBSe-PEG的熔融温度(Tm)和结晶温度(Tc)逐渐下降;相比PBS,脂肪酶N435对PBSe-PEG具有更好的降解性。  相似文献   

12.
采用“一步法”,以丁二酸酐(SAA)和1,4–丁二醇(BDO)为单体、端羟基二元醇为共聚单体合成了聚丁二酸丁二醇酯(PBS)及一系列端羟基二元醇共聚物,同时使酚酞与SAA的缩聚产物参与SAA和BDO的共聚反应,并通过链段调节合成法制备兼具刚性链段和柔性链段的可生物降解三嵌段共聚聚酯热塑性弹性体聚(丁二酸丁二醇酯-共-酚酞丁二酸丁二醇酯)(SAA-PHE-PBS),研究了PBS及其共聚物的分子量、化学结构组成、热性能和结晶性能,此外,使用南极假丝酵母脂肪酶B测试了PBS及其共聚物的生物降解性能。结果表明,端羟基二元醇共聚物的玻璃化转变温度变化幅度不大,熔融温度无明显改变,结晶度降低,亲水性有所改善,生物降解性能得到大幅度提升;三嵌段热塑性弹性体SAA-PHE-PBS的玻璃化转变温度升高,结晶度与PBS相差不大,疏水性更强,共聚合物的残重率有所增加,生物降解性能有不同程度的降低。  相似文献   

13.
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   

14.
(丁二酸丁二酯/丁二酸己二酯)共聚物的合成及性能   总被引:1,自引:0,他引:1  
以丁二酸、丁二醇、己二醇为原料,在十氢萘中进行直接缩聚反应,合成了高分子量(丁二酸丁二酯/丁二酸己二酯)共聚物,产率达到95%以上。FT—IR和^1H—NMR图谱表明,共聚物的结构为预期结构;GPC测试结果表明,共聚物均具有较高的分子量;与聚丁二酸丁二酯(PBS)相比,共聚物的拉伸强度显著降低,但断裂伸长率有所提高:DSC测试结果表明,共聚物的结晶度明显低于PBS,其熔点、结晶温度随体系中丁二酸己二酯单元的增加而降低:TG测试结果表明,共聚物均具有较好的热稳定性。  相似文献   

15.
Copolymers of ε-CL/L-LA and ε-CL/DL-LA were allowed to age in a buffer solution of pH 7 at 23 and 37°C. The effects of time and temperature on the rate of hydrolysis were examined by various techniques including weighing (water absorption and weight loss), SEC (molecular weight and polydispersity), and DSC (thermal properties). For comparison, the hydrolytic behavior of PLLA, PDLLA, and commercial PCL homopolymers was investigated by the same methods. SEC measurements showed that molecular weights of the copolymers and PLA homopolymers started to decrease during the first week of hydrolysis, but significant mass losses occurred only much later. As expected, there was no change in either molecular weight or mass of PCL during the hydrolysis study. The kinetic results for copolymers and homopolymers were calculated to study the degradation mechanism. During hydrolysis, the crystallinity of the initially semicrystalline copolymers increased and some crystallinity appeared in the initially amorphous L-LA-containing copolymers. © 1996 John Wiley & Sons, Inc.  相似文献   

16.
The degradation behavior of P(TMC-co-CL) in different compositions was investigated via subcutaneous implantation in vivo. To clarify the role of enzymes in the degradation behavior of the copolymers, hydrolytic and enzymatic degradation were also performed. The mass loss, changes in molecular weight and polydispersity, as well as the variation in composition were monitored with degradation. The changes in thermal and mechanical properties of the specimens were also studied. The results showed that the preferred cleavage of ester bonds resulted in faster degradation in both the hydrolytic and enzymatic cases. Furthermore, the P(TMC-co-CL) had a higher degradation rate in the presence of lipase because it cleaves ester bonds as well as the role of surfactants in the diffusion of the degradation products into water. In vivo, the degradation behavior of the P(TMC-co-CL) depended on their composition—copolymers with a higher TMC content degraded primarily via surface erosion. Bulk degradation was observed for those with a higher CL content. After degradation the mechanical properties and thermal stabilities of the copolymers deteriorated, but the Tm and crystallinity increased via preferred degradation of the amorphous regions. The P(TMC-co-CL) had a tunable degradation rate and remains a promising candidate for clinical subcutaneous implants especially through form-stabilization work.  相似文献   

17.
倪洪凯  杨彪  许国志  李明 《塑料工业》2012,40(1):64-67,81
采用湿法共混工艺制备了聚丁二酸/己二酸–丁二醇酯(PBSA)/木薯淀粉薄膜。对薄膜的力学性能、热稳定性和微观形貌以及树脂的摩尔质量变化进行了研究。结果表明,共混体系的熔融温度基本没有变化,结晶峰温度略有升高;随木薯淀粉含量的增加,PBSA/淀粉共混薄膜的力学性能下降,木薯淀粉和PBSA在各自的温度区域内分解,共混材料的热稳定性下降;GPC结果显示加工过程中,虽有水分存在,但是PBSA的摩尔质量没有降低。随淀粉含量的增加,薄膜的拉伸强度和断裂伸长率逐渐降低,当淀粉质量分数10%时,共混薄膜仍能保持良好的机械性能,达到GB/T 4456—2008的要求。湿法共混工艺能够在一定程度上取得物料共混的理想效果,在降低生产成本的同时保持良好的综合机械性能和加工性能。  相似文献   

18.
A series of thermoplastic elastomers based on ethylene oxide‐poly(dimethylsiloxane)‐ethylene oxide (EO‐PDMS‐EO), as the soft segment, and poly(butylene terephthalate) (PBT), as the hard segment, were synthesized by catalyzed two‐step, melt transesterification reaction of dimethyl terephthalate (DMT) with 1,4‐butanediol (BD) and α,ω‐dihydroxy‐(EO‐PDMS‐EO). Copolymers with a content of hard PBT segments between 40 and 90 mass % and a constant length of the soft EO‐PDMS‐EO segments were prepared. The siloxane prepolymer with hydrophilic terminal EO units was used to improve the miscibility between the polar comonomers, DMT and BD, and the nonpolar PDMS. The molecular structure and composition of the copolymers were determined by 1H‐NMR spectroscopy, whereas the effectiveness of the incorporation of α,ω‐dihydroxy‐(EO‐PDMS‐EO) into the copolymer chains was verified by chloroform extraction. The effects of the structure and composition of the copolymers on the melting temperatures and the degree of crystallinity, as well as on the thermal degradation stability and some rheological properties, were studied. It was demonstrated that the degree of crystallinity, the melting and crystallization temperatures of the copolymers increased with increasing mass fraction of the PBT segments. The thermal stability of the copolymers was lower than that of PBT homopolymer, because of the presence of thermoliable ether bonds in the soft segments. © 2009 Wiley Periodicals, Inc. J Appl Polym Sci, 2010  相似文献   

19.
The catalytic degradation of prepared matrix poly(butylene succinate) (PBS) and its copolymers by immobilized lipase is carried out in the mixed organic solvent containing a small amount of water. The degradation products were studied with various characterizations techniques, including gel permeation chromatography (GPC), time of flight mass spectrum (TOF‐MS), nuclear magnetic resonance (NMR), and Fourier transform‐infrared spectroscopy (FT‐IR). The results showed that under atmospheric pressure, 60°C, after 24 h catalytic degradation of PBS and its copolymers by immobilized lipase, the yellow oil‐like degradation products can be obtained. The lipase has catalytic activity on various copolymers. At the first time, the monomer of BDO was found in the degradation products and the molecular weight of product with aromatic smell is below 1000. The products consisted of cyclic oligomer, linear oligomer and monomers, and cyclic oligomer is at least dimmer. The minimum and maximum degradation yields correspond to PBS (40%) and P(BS‐co‐CL‐co‐CHDM) (54%). © 2011 Wiley Periodicals, Inc. J Appl Polym Sci, 2012  相似文献   

20.
A new series of biodegradable aliphatic–aromatic nanocomposites containing maleic acid‐grafted poly[(butylene adipate)‐co‐terephthalate] (g‐PBAT) and organically modified layered zinc phenylphosphonate (m‐PPZn) were successfully synthesized through transesterification and polycondensation processes with covalent linkages between the polymeric and inorganic materials. Fourier transform infrared and 13C NMR spectra demonstrate the successful grafting of maleic acid to PBAT. The morphology of g‐PBAT/m‐PPZn nanocomposites was investigated using wide‐angle X‐ray diffraction and transmission electron microscopy. Results showed that the stacking layers of m‐PPZn were distributed and intercalated into the g‐PBAT polymer matrix. The incorporation of m‐PPZn into the g‐PBAT matrix significantly enhanced the storage modulus at ?70 °C as compared to that of neat g‐PBAT. A reduction in thermal stability was observed for all g‐PBAT/m‐PPZn systems, which is probably due to the lower thermal stability of m‐PPZn. The biodegradation of neat g‐PBAT copolymers and g‐PBAT/m‐PPZn nanocomposites was investigated using lipase from Pseudomonas sp. The degradation rates of neat g‐PBAT copolymers decrease in the order g‐PBAT‐80 > g‐PBAT‐50 > g‐PBAT‐20. The faster degradation rate of g‐PBAT‐80 is a result of the higher content of adipate acid units and the chain flexibility of the polymer backbone. Furthermore, the weight loss increases as the loading of m‐PPZn increases, indicating that the presence of m‐PPZn improves the degradation of the g‐PBAT copolymers. This result might be accounted for by the lower degree of crystallinity for g‐PBAT/m‐PPZn nanocomposites. © 2019 Society of Chemical Industry  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