首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到19条相似文献,搜索用时 812 毫秒
1.
用癸二酸、己二酸部分取代丁二酸,通过熔融溶液相结合的聚合方法,得到两者一系列的共聚物。通过数均分子量和熔点两个指标,得出最佳取代比例为20%癸二酸取代丁二酸的共聚物(PBSE-20)和20%己二酸取代丁二酸的共聚物(PBSA-20)。用异佛尔酮二异氰酸酯(IPDI)对这两种共聚物进行扩链改性,结果表明:扩链后,其数均分子量大小顺序为PBS>PBSA-20>PBSE-20,分子量分布均小于扩链前;且断裂伸长率均大于扩链前,各聚合物的热分解温度均大于300℃。阐述了脂肪族聚酯分子量、熔点及分子结构对称性与降解性能之间的关系。  相似文献   

2.
脂肪族聚酯的合成及降解性能研究   总被引:1,自引:1,他引:0  
通过熔融和溶液结合法合成了脂肪族聚酯—聚丁二酸丁二醇酯(PBS)、丁二酸丁二醇酯-己二酸丁二醇酯共聚物(P(BS-co-BA))和丁二酸丁二醇酯-癸二酸丁二醇酯共聚物(P(BS-co-BSe)),并对其进行了生物降解实验。采用GPC测定了脂肪族聚酯的分子量及其分布,并采用熔点仪测定了聚酯的熔点。得到脂肪族聚酯分子量、熔点以及分子结构对称性和降解之间的关系。  相似文献   

3.
寇莹 《广州化工》2012,(4):39-41
为了探讨线性PBS基脂肪族共聚酯的结构和降解之间的关系,首先合成了线性PBS基脂肪族共聚酯,即聚丁二酸丁二醇酯-共-聚己二酸丁二醇酯P(BS-co-BA),聚丁二酸丁二醇酯-共-癸二酸丁二醇酯P(BS-co-BSe),并将线性PBS基脂肪族共聚酯及PBS在土壤悬浮液中进行降解,通过GPC、熔点测定仪对线性PBS基脂肪族共聚酯的分子量和熔点进行了测定;通过测定降解过程中失重率和降解前后聚酯薄膜表面形貌来对共聚酯降解程度进行表征。结果表明:随着二元酸碳链的增长,分子对称性降低,降解性能增大。通过观察分子量,熔点及降解失重率的测定结果,得出分子量越大,降解越不容易进行;熔点越小,降解性能越好。  相似文献   

4.
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   

5.
以1,4-丁二醇及不同比例的己二酸、丁二酸为原料,制备一系列聚丁二酸-己二酸丁二醇酯[P(BS-co-BA)]共聚酯,借助衰减全反射傅里叶变换红外光谱仪(ATR-FTIR)、核磁共振仪(1H-NMR)、差示扫描量热仪(DSC)、X射线衍射仪(XRD)以及热重分析仪(TG)等对共聚酯的结构性能进行表征分析。角质酶降解结果表明,在经过16 h后,4种共聚酯降解率均达到80%以上,2种均聚酯均仅为40%左右。其中聚酯酶降解速率为P(BS-co-40%BA)>P(BS-co-60%BA)>P(BS-co-80%BA)>P(BS-co-20%BA)>聚己二酸丁二醇酯(PBA)>聚丁二酸丁二醇酯(PBS)。综合可知P(BS-co-40%BA)热稳定性相比于PBA更好,降解性能较PBS更好,为最佳共聚酯。  相似文献   

6.
张敏  张若琳  荆晶晶  王蕾  邱建辉 《塑料》2014,(3):48-50,80
通过控制反应条件的方法,以己二酸(BA)、乙二醇(EG)、己二醇(HS)作为改性PBS的共聚单体,合成了几种低分子质量的PBS基共聚物。采用1H-NMR、FT-IR表征了共聚物的化学结构;WXRD、TGA、DSC对共聚物的性能进行了研究。结果表明:酯化时间为2 h,抽真空时间在30 min以内时,能够得到数均分子质量5 000以上的PBS基共聚物。对于相近分子质量的不同共聚物,聚合后结晶形态没有明显变化,但结晶度和熔点都有所下降;相比PBS,P(BS-co-BA)和P(BS-co-EG)的热分解温度都有所增加,而P(BS-co-HS)的有所降低,但分解温度仍在300℃以上。  相似文献   

7.
PBS不同化学结构共聚物的性能   总被引:5,自引:2,他引:3  
采用改变原料的组分合成不同化学结构的聚丁二酸丁二醇酯(PBS)改性共聚物:聚丁二酸丁二醇/己二酸丁二醇酯(PBS-co-BA)、聚丁二酸丁二醇酯/聚丁二酸乙二醇酯(PBS-co-ES)、聚丁二酸丁二醇酯/聚丁二酸己二醇酯(PBS-co-HS),利用FT-IR和1H-NMR表征共聚物的化学结构,并对共聚物的结构与物理性能、降解性能的关系进行对比.研究结果表明:所有共聚物的结晶度、熔点较其均聚物有所降低;但所有共聚物的断裂伸长率都有所提高.热分析结果表明:PBS-co-HS热性能有所提高,PBS-co-BA和PBS-co-ES有所下降.堆肥降解实验表明:所有共聚物的降解性都比均聚物有显著提高,其降解速度大小顺序为:PBS-co-BA>PBS-co-HS>PBS-co-ES>PBS,PBS-co-HS是综合性能最优良的材料.  相似文献   

8.
将扩链剂与聚丁二酸丁二醇酯(PBS)共混,制备改性PBS薄膜。通过万能拉伸试验机、差示扫描量热仪(DSC)、傅里叶变换红外光谱仪(FTIR)和旋转流变仪等,表征PBS薄膜和扩链改性PBS薄膜室温降解老化后结构与性能的变化。结果表明:室温老化降解4周后,PBS薄膜的应变100%拉伸强度较改性PBS薄膜大幅度下降,扩链剂改性PBS可以减缓拉伸性能老化衰减。随着老化时间增加,改性前后PBS薄膜在室温老化过程储能模量、损耗模量、复数黏度明显下降,未改性PBS薄膜变化更明显。PBS薄膜在室温下降解未对晶区造成破坏,主要发生无定形区。而扩链改性使PBS分子链增加,减少PBS膜非晶区分子链断裂,有助于抑制无定型区降解老化。  相似文献   

9.
邹俊  魏芸  吴毅炳  张竞  李世云 《塑料工业》2013,41(5):26-29,42
以丁二酸和1,4-丁二醇为原料,采用熔融缩聚法合成了聚丁二酸丁二醇酯(PBS)预聚物,再与L-丙交酯(L-LA)开环共聚,合成聚乳酸/聚丁二酸丁二醇酯嵌段共聚物(PLLA-co-PBS)。研究了共聚物的结构、热性能、结晶性能和亲水性。结果表明,PBS与L-LA开环共聚生成了PLLA-co-PBS嵌段共聚物;PLLA-co-PBS嵌段共聚物经两个阶段的热分解,且PBS链段的引入提高了聚合物的热稳定性;随着PBS引入量的增加,聚合物的结晶性能,亲水性能都有一定的提高。  相似文献   

10.
通过熔融溶液相结合的方法合成聚丁二酸丁二醇酯(PBS)预聚物。用异佛尔酮二异氰酸酯(IPDI)、六亚甲基-1,6-二异氰酸酯(HDI)和丁二酰氯(SUC)对PBS预聚体进行扩链改性。采用熔点仪测定PBS的熔点;采用傅里叶变换红外光谱仪对PBS的结构进行了表征;采用凝胶渗透色谱仪、热重仪和万能实验机测定了PBS的摩尔质量及其分布、热稳定性和力学性能;测定失重率,考察不同扩链剂对PBS降解性能的影响。结果表明:PBS扩链后摩尔质量均有显著提高,摩尔质量分布均小于扩链前;且断裂伸长率等力学性能和热分解温度等热性能也都有明显改善。但用不同扩链剂扩链后,PBS降解性能有较大区别。  相似文献   

11.
王凯  厉安昕 《辽宁化工》2011,40(6):549-551,560
采用扩链剂2,4-甲苯二异氰酸酯(TDI)、2,2′-双(2-恶唑啉)(BOZ)以及新型扩链剂ADR-4370对聚丁二酸丁二醇酯(PBS)进行扩链,对比研究3种扩链剂对PBS性能的影响。采用聚乳酸(PLA)与PBS共混,选择最佳配比以及合适的共混工艺。结果表明,新型扩链剂ADR-4370可以良好的改善PBS的加工性能和机械性能,m(PBS)/m(PLA)最佳配比为70:30,在共混同时进行扩链比扩链后再共混可以更好的提高PBS/PLA的性能。  相似文献   

12.
以聚丁二酸丁二酯(PBS)薄片为降解底物,利用角质酶对其进行降解研究,考察PBS的酶促降解行为。利用扫描电子显微镜、差示扫描量热仪和热重分析仪等方法对降解前后的PBS薄片进行了表征分析,并进一步采用质谱仪对降解产物进行分析。结果表明,在酶浓度2.5 U/mL、反应温度37 ℃以及pH 7.4的条件下,经16 h降解PBS薄片的降解率可达93.88 %,在降解时间为2 h时有最大降解速率32.97 μg/cm2·h;PBS薄片有片层脱落降解的现象,促进了角质酶的降解作用;随降解时间的延长,PBS相对结晶度逐渐降低,热稳性也呈现下降趋势;PBS被降解成单体或寡聚物。  相似文献   

13.
This study investigates the influence of three fatty acids (lauric acid, palmitic acid, and stearic acid) on biodegradable polymer blends based on poly(lactic acid) (PLA) and poly(butylene succinate) (PBS), containing different weight ratios (100:0, 100:2, and 100:4) of fatty acids on the transparency, mechanical properties, morphology, contact angle, and water vapour permeability. All of the blends were pressed into thin films and tested. The experimental results showed that the properties of the samples varied with chain length and amounts of the fatty acids. Thus, it could be concluded that use of fatty acids opens up new ways for the plasticisation of PLA/PBS blends for use as new bioplastics.  相似文献   

14.
The effect of hydrolytic degradation on the microstructure of unoriented, random 50: 50 poly(glycolic acid-co-D ,L -lactic acid) was examined using simultaneous small and wide angle X-ray scattering (SAXS/WAXS) and differential scanning calorimetry (DSC). Samples were degraded in phosphate-buffered saline solution at 37·5°C and studied wet and after dehydration. There was no evidence of crystalline material within the sample at any stage of degradation or dehydration from either X-ray scattering or thermal analysis. Thus, chain scission does not enable crystallization of the copolymer, and the glycolic acid and lactic acid fragments formed on degradation do not crystallize, even when the samples are dehydrated. Because such fragments are clearly formed (Hakkarainen, M., Albertsson, A. C. & Karlsson, S., Polym. Deg. Stab., 52 (1996) 283), and because they are crystalline in the dry state, it must be assumed either that these species are not present in any quantity in the degrading sample and that they diffuse easily from the bulk into the surrounding medium, or that the bulk polymer prevents them from crystallizing. SAXS gave evidence of small voids within the structure. Unlike dehydrated degraded semi-crystalline samples, there is no evidence for voiding on a macroscopic scale. The number and size of the small voids in the dehydrated samples rises with degradation. The voids close as samples are heated above the glass transition temperature and the amorphous chains gain mobility. The glass transition, although clearly visible in the undegraded samples, becomes less visible by DSC on degradation. After 28 days’ degradation, there is some evidence that the structure begins to close up, perhaps as a result of reduced viscosity arising from the increased fraction of low molecular weight material. © 1998 SCI.  相似文献   

15.
The ultrasonic degradation of poly(acrylic acid), a water‐soluble polymer, was studied in the presence of persulfates at different temperatures in binary solvent mixtures of methanol and water. The degraded samples were analyzed by gel permeation chromatography for the time evolution of the molecular weight distributions. A continuous distribution kinetics model based on midpoint chain scission was developed, and the degradation rate coefficients were determined. The decline in the rate of degradation of poly(acrylic acid) with increasing temperature and with an increment in the methanol content in the binary solvent mixture of methanol and water was attributed to the increased vapor pressure of the solutions. The experimental data showed an augmentation of the degradation rate of the polymer with increasing oxidizing agent (persulfate) concentrations. Different concentrations of three persulfates—potassium persulfate, ammonium persulfate, and sodium persulfate—were used. It was found that the ratio of the polymer degradation rate coefficient to the dissociation rate constant of the persulfate was constant. This implies that the ultrasonic degradation rate of poly(acrylic acid) can be determined a priori in the presence of any initiator. © 2009 Wiley Periodicals, Inc. J Appl Polym Sci, 2009  相似文献   

16.
Mechanical degradation and mechanochemical reaction in heterogeneous and homogeneous systems of poly(vinyl chloride) and poly(ethylene-co-propylene) polymer have been studied by ultrasonic irradiation at 30 °C. The rates of decrease in the number-average molecular weights of the degraded poly(vinyl chloride) and poly(ethylene-co-propylene) polymer were much faster in order of the solid poly(vinyl chloride)—poly(ethylene-co-propylene) polymer solution, the swelled poly(vinyl chloride)—poly(ethylene-co-propylene) polymer solution, and the homogeneous solution systems. On the other hand, mechanochemical reaction occurred by free radicals produced from the chain scissions of both polymers by ultrasonic waves. The changes in the composition of the total block copolymer, the unreacted poly(vinyl chloride), and the unreacted poly(ethylene-co-propylene) polymer in individual reaction systems were obtained. In addition, the microscopic observation of the surfaces of the polymers on before and after mechanochemical reaction is carried out. Received: 10 May 2000/Revised version: 1 August 2000/Accepted: 3 August 2000  相似文献   

17.
A poly(3‐hydroxybutyrate‐co‐4‐hydroxybutyrate) (P3/4HB)‐degrading strain, Agrobacterium sp. DSGZ, was isolated from sewage by poly(3‐hydroxybutyrate) (PHB) mineral agar plates. A novel P3/4HB depolymerase with a molecular weight of 34 kDa was purified through a novel single‐step affinity chromatography method from the culture supernatant of the strain by using P3/4HB powder as a substrate. The purified depolymerase showed optimum activity at pH 7.0 and 50°C, and was stable at the pH range of 6.0 to 9.0 and temperature below 50°C. Enzyme activity was strongly inhibited by phenylmethylsulfonyl fluoride (PMSF), ethylenediaminetetraacetic acid (EDTA), hydrophobic reagents, and some metal ions. The depolymerase degraded poly(3‐hydroxybutyrate) (PHB), poly(hydroxybutyrate‐co‐hydroxyvalerate) (PHBV), P3/4HB, and polycaprolactone (PCL), instead of polylactic acid (PLA) or poly(butylene succinate) (PBS). Meanwhile, the depolymerase showed high hydrolytic activity against short‐chain length esters, such as butyrate acid ester and caprylic acid ester. The main degradation products of the depolymerase were identified as hydroxybutyrate monomers and dimers, and the monomers were identified as 3‐hydroxybutyrate (3HB) monomers and 4‐hydroxybutyrate (4HB) monomers. The preparation procedure, crystallinity, and 4HB composition of the P3/4HB copolymer showed evident effect on degradation behavior, and change in crystallinity was the main factor affecting degradation. © 2015 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2016 , 133, 42805.  相似文献   

18.
In order to improve the viscoelasticity and foamability, poly (butylene succinate) (PBS) was modified through reactive melt mixing with chain extender (CE) having multi epoxy‐groups. Subsequently, the prepared chain extended PBS (CEPBS) was foamed in a high pressure stainless steel autoclave using CO2 as physical blowing agent. The molecular weight, thermal properties, rheological properties, and foam properties of various PBS samples were characterized using gel permeation chromatography, differential scanning calorimetry, rotational rheometer, and scanning electron microscope, respectively. With the introduction of CE, the molecular weight, the crystallization temperature, the complex viscosity and storage modulus of PBS were increased and the crystallization degree of PBS was decreased. At the CE content of 0.75 phr, the cross‐linking structure was formed and the expansion volume ratio increased to nearly 15 times, which meant the chain extension played an important role in the foaming process of PBS. POLYM. ENG. SCI., 55:988–994, 2015. © 2014 Society of Plastics Engineers  相似文献   

19.
Poly(butylene terephthalate)‐co‐poly(butylene succinate)‐block‐poly(ethylene glycol) segmented random copolymers, with poly(butylene succinate) (PBS) molar fraction (MPBS) varying from 10 to 60 %, were synthesized through a melt polycondensation process and characterized by means of GPC, NMR, DSC and mechanical testing. The number‐average relative molecular mass of the copolymers was higher than 4 × 104 g mol?1 with polydispersity below 1.9. Sequence distribution analysis on the two types of hard segments by means of 1H NMR revealed that the number‐average sequence length of PBT decreased from 2.80 to 1.23, while that of PBS increased from 1.27 to 4.76 with increasing MPBS. The random distribution of hard segments was also justified because of the degree of randomness around 1.0. Micro‐phase separation structure was verified for the appearance of two glass transition temperatures and two melting points, respectively, in DSC thermograms of most samples. The crystallinity of hard segments changed with the crystallizability controlled by the average sequence length and reached the minimum value at an MPBS of about 50–60 mol%. The results can also be ascribed to the co‐crystallization between two structurally analogous hard segments. Mechanical testing results demonstrated that incorporating a certain amount of PBS moieties (less than 30 mol%), at the expense of a minute depression of the elastic modulus, that higher relative elongation and more flexibility of polymer chain could be expected. Maximum equilibrium water absorption and faster degradation rates were observed on samples with higher MPBS values and lower crystallinity of hard segments were better hydrophilicity of the polymer chain, through in vitro degradation experiments. Copyright © 2003 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