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 218 毫秒
1.
叶斌  高才  刘向农  杨锁  江斌 《物理化学学报》2011,27(5):1031-1038
采用差示扫描量热法(DSC)测定山梨醇样品经历不同时间(ta)等温退火后, 以10 K·min-1速率进行升温时玻璃化转变温度(Tg)前后的比热容(Cp(T)). 将Gómez Ribelles (GR)提出的一种基于构型熵的现象学模型用于描述山梨醇玻璃的焓松弛行为, 考察GR模型能否适用于小分子玻璃体系. 结果表明, 单组GR模型参数拟合的曲线均能较好重现对应热历史条件下的山梨醇体系的实验所得Cp(T)曲线, 尽管并未找到不随热历史而变的一组参数作为材料常数, 但与其它现象学模型应用于小分子玻璃时, 其模型参数都随热历史变化而变化的特点相比, GR模型的某些参数基本保持不变. 且在较长退火时间下拟合得到的模型参数普适性较好. 同经历连续降温的山梨醇相比, 等温退火过程得到的松弛极限态参数(δ)的平均值与Tg处比热容增量(ΔCp(Tg))的比值明显增大, 但仍小于聚合物的值, 表明GR模型提出的亚稳极限态对小分子玻璃的影响值得商榷.  相似文献   

2.
高才  王铁军  周国燕  华泽钊 《化学学报》2007,65(21):2393-2400
为了验证基于位形熵的非线性Adam-Gibbs协同松弛模型(AGV)能否用于描述小分子氢键液体的协同松弛行为, 利用差示扫描量热法(DSC)测量了连续升降温条件下1,2-丙二醇及其四种水溶液在115~230 K之间的比热容, 利用曲线拟合技术获得AGV模型参数. 结果表明, AGV模型可以重现体系的实验比热容数据. 1,2-丙二醇表现出与其水溶液明显不同的松弛行为, 但水含量的变化对松弛行为的影响并不明显. 利用AGV方法和Johari方法分别对协同重排活化能(Δμ')和协同重排域(CRR)尺寸(z*)作了分析. 只有选择比聚合物大得多的某一协同重排位形数, 以AGV方法得到的z*才不至于没有物理意义. Johari方法的分析结果表明, Tg温度下1,2-丙二醇的CRR内有约3个分子, 但对应的协同重排位形数(W*)却较聚合物高出很多. Donth的基于热力学温度波动理论的分析表明, 1,2-丙二醇及其水溶液的CRR尺寸随组分的变化趋势可与Δμ'和非指数参数的分析结果相吻合, 但得到的1,2-丙二醇的CRR内有约350个分子, 从而和Johari的分析结果产生巨大差别.  相似文献   

3.
为了考察木糖醇的玻璃化转变和焓松弛行为,寻求碳链长度对线性多元醇玻璃化转变和焓松弛行为的影响,利用差示扫描量热(DSC)技术测定了不同降温速率下木糖醇在玻璃化转变温度(Tg)前后的比热容(Cp),通过曲线拟合获得了TNM(Tool-Narayanaswamy-Moynihan)模型参数,并和其他多元醇类已有研究结果进行对照.结果表明,尽管TNM模型可以很好地重现不同降温速率体系的实验比热容数据,但模型参数并不是材料常数,而是和热历史有关,不同的降温速率对应不同的模型参数.指前因子(A)、非线性参数(x)和非指数参数(β)均随着降温速率的增加而降低,松弛活化焓(△h*)的变化趋势刚好相反.几种线性多元醇玻璃化转变和TNM模型参数的对照表明,玻璃化转变温度,松弛活化焓和动力学脆度(m)都随着烷基碳链长度的增加而增加.虽然非线性参数、非指数参数随碳链长度的增加有降低的趋势,但木糖醇展现出反常变化的情形.  相似文献   

4.
曹晨忠  高硕 《化学学报》2007,65(24):2898-2904
将芳环上取代基的电子效应参数引入卤代甲烷, 以卤代甲烷分子Y-CHnX3-n (n=0~3; Y=H, F, Cl, Br, I; X=F, Cl, Br, I)中Y-C键的标准键焓 与中心C原子相键连原子的场/诱导效应之和ΣFi、共轭效应之和ΣRi以及诱导偶极之和Σ(α×F)为参数, 建立了一个定量估算卤代甲烷分子中Y-C键离解能(BDE)的通用模型, BDE(Y-C)=57.5460+0.8855 -101.0780ΣRi-64.8390ΣFi-10.1034Σ(α×F). 对35个C-H, C-F, C-Cl, C-Br和C-I键回归分析结果表明, 估算Y-C键离解能的精度在实验误差范围内. 对外部数据集的预测结果表明, 该模型具有较高的预测精度, 可用于预测还没被实验测定的卤甲烷中Y-C键离解能. 还对卤代甲烷中104个C-Y键的键离解能进行了预测. 将芳环上取代基效应用于研究饱和体系化学键性能, 有利于深入理解取代基效应对化学键性能的影响.  相似文献   

5.
采用等温滴定量热法研究植物凝集素(PHA)作用于20%家兔血红细胞(RBC)的凝集反应过程, 以生物热动力学参数表达凝集反应动力学特征. 结果表明, PHA与家兔血红细胞凝集反应的平衡常数K为3.24×105 L•mol-1, 反应焓变( )为-238.548 kJ•mol-1, 反应自由能( )为-32.722 kJ•mol-1, 反应熵变( )为-0.664 kJ•mol-1•K-1. 反应为放热反应(Q>0), 可自发进行(ΔG<0), 反应为焓驱动、熵减过程(ΔH<0, ΔS<0, |ΔH|>|TΔS|), 由化学键合作用、氢键和范德华力推动, 反应体系最终处于稳定有序状态. 所建立的微量量热法可以客观表征凝集反应的热动力学过程, 也为其他抗原抗体类型反应的过程评价方法提供新的技术参考.  相似文献   

6.
王晓蜂  袁荞龙 《化学学报》2012,70(9):1047-1054
以(N,N-二甲氨基-4-吡啶)五氰合铁(II)封端的聚氧丙烯聚氧乙烯共聚物(EPE-Fe)与苯乙烯在水中自组装形成纳米体系(EPE-Fe-St), 在纳米尺度受限空间内进行了苯乙烯自由基聚合, 制备了聚苯乙烯微球(EPE-Fe-PS). 用Fe3+对自组装体系的纳米球壳进行固化后形成Fe-EPE-Fe-St 体系, 聚合后也制备了聚苯乙烯微球(Fe-EPE-Fe-PS). 研究结果表明,制备了粒径为60~200 nm 的不同粒径单分散聚苯乙烯微球, 聚合温度对纳米Fe-EPE-Fe-St 体系粒径影响较小, 而对EPE-Fe-St 体系较大. 在受限空间内苯乙烯的自由基聚合可得到数均分子量超过70 万的聚苯乙烯; 自组装体系中引发剂量增多使聚苯乙烯分子量下降, 聚合温度上升也使分子量下降, 而增加自组装的EPE-Fe 用量可增加聚苯乙烯的分子量. 两种受限条件下的聚苯乙烯微球的玻璃化转变温度(Tg)在90~135 ℃之间, 纳米反应器壳层的硬化提高了聚苯乙烯微球的Tg.  相似文献   

7.
纳米零价铁(Nano zero-valent iron,nZVI)被广泛应用于水污染治理,高纯度且分散性良好的nZVI的制备方法一直是研究热点.本文采用含不同羟基数目的醇(乙醇、乙二醇、赤藓糖醇、甘露醇和山梨醇)作为改性剂,分别制备得到n ZVI-EA,nZVI-EG,nZVI-ER,nZVI-M和nZVI-S样品.将上述样品应用于水中微囊藻毒素(Microcystin-LR,MC-LR)的还原去除.结果表明,随着改性剂羟基数目的增多,改性nZVI的抗氧化能力和分散性增强,对MC-LR的降解反应速率也随之提高.nZVI-M去除MC-LR的表面积校正特征速率常数(79.35×10-5L·m?2·min?1)是nZVI-S(8.55×10-5 L·m?2·min?1)的9.3倍,是未改性样品nZVI0(1.30×10-5 L·m?2·min?1)的61.0倍.通过X射线衍射...  相似文献   

8.
咸德玲  黄可龙  刘素琴  肖静怡 《化学学报》2007,65(23):2663-2668
脂质体电动色谱是一种理想的评价药物与生物膜相互作用的模型. 在288~323 K范围内测定了中性芳香族溶质在脂质体电动色谱中的分配系数, 通过三项式拟合Van't Hoff图获得了一系列的热力学参数, 研究了溶质在脂质体电动色谱中的热力学分配行为. 结果表明, 分配系数随体系温度的升高和苯环所带亚甲基数目的增加而增大. 从288到323 K, ΔH>0, -TΔS<0, ΔG<0, 溶质在脂质体电动色谱中的分配过程为熵驱动过程. 从288到298 K, 脂质体电动色谱分配系统的ΔCp为负值, 其表现行为与经典疏水作用一致. 从303到323 K, 脂质体电动色谱分配系统的ΔCp为正值, 其表现行为与经典疏水作用不完全吻合. ΔH和ΔS呈线性关系, 该分配系统存在焓熵补偿.  相似文献   

9.
王军  杨许召  武金超  宋浩  邹文苑 《色谱》2015,33(12):1301-1306
在343.15~363.15 K下采用反气相色谱法对实验室合成的3种非对称双阳离子型离子液体[PyC5Pi]NTf2]2、[MpC5Pi]NTf2]2和[PyC6Pi]NTf2]2的溶解度参数进行了测定。以正辛烷、正癸烷、正十二烷、正十四烷和正十六烷作为探针溶剂,计算了在不同温度下探针溶剂在3种离子液体中的特性保留体积(Vg0)、摩尔吸收焓(ΔH1S)、无限稀释摩尔混合焓(ΔHl)、摩尔蒸发焓(ΔHv)、无限稀释活度系数(Ω1)以及探针溶剂与3种离子液体的Flory-HuggirIs相互作用参数(χ12),得到了室温下3种离子液体的溶解度参数(δ2)为28.52~32.66 (J·cm-3)1/2。分子结构中含有4-甲基吗啉比含有吡啶时溶解度参数更大。随着两个阳离子间连接基碳数的增加,溶解度参数增大。这一结果对研究离子液体的溶液性质和应用有指导作用。  相似文献   

10.
王晓星  顾彦  陈登霞  方艳芬  黄应平 《化学学报》2010,68(23):2463-2470
以小牛胸腺DNA为对象, 在超氧化物歧化酶(superoxide dismutase, SOD)存在下, 研究了纳米二氧化钛(TiO2)光催化对DNA损伤特性及SOD对TiO2光催化损伤DNA的影响. 采用凝胶电泳和高效液相色谱(HPLC)分析法, 探讨了DNA的损伤程度. 运用生物标准样8-羟基脱氧鸟苷(8-hydroxy-2 -deoxyguanosine, 8-OHdG)为内标物并通过高效液相色谱-电喷雾离子化串联质谱联用技术(HPLC-ESI-MS/MS)对DNA损伤产物进行了跟踪分析, 采用过氧化物酶催化分光光度法及顺磁共振技术(ESR)跟踪测定TiO2光催化DNA损伤过程中H2O2和氧化物种的变化, 探讨了DNA氧化损伤机理. 结果表明, 在实验条件下紫外光照(UV, λ=200~275 nm), Dark/TiO2和UV/TiO2体系中, DNA氧化损伤程度为UV/TiO2>UV>Dark/TiO2. 光催化260 min DNA损伤99%, 反应动力学常数K=7.82×10-3 min-1. 抗氧化剂SOD具有清除光催化体系中超氧自由基( )的能力, 可以抑制DNA的损伤, 反应动力学常数K=2.27×10-3 min-1. 8-OHdG为DNA损伤中鸟嘌呤氧化的特异产物, UV及光催化体系对DNA损伤主要涉及 及羟基自由基(•OH)历程, 光催化体系对DNA损伤伴随有深度氧化(矿化)过程, 实验条件下12 h DNA矿化75.06%.  相似文献   

11.
The structural relaxation process in styrene-acrylonitrile copolymer has been characterized by means of differential scanning calorimetry (DSC) experiments. The results in the form of heat capacity, cp(T), curves are analyzed using a model for the evolution of the configurational entropy during the process recently proposed by the authors.11,12 The model simulation allows one to determine the enthalpy (or entropy) structural relaxation times and the β parameter of the Kohlrausch-Williams-Watts equation characterizing the width of the distribution of relaxation times. This material parameters are compared with their analogues determined from the dielectric and dynamic-mechanical relaxation processes. © 1997 John Wiley & Sons, Inc. J Polym Sci B: Polym Phys 35: 2201–2217, 1997  相似文献   

12.
Ethylene-propylene (EP) and ethylene-octene (EO) copolymers polymerized with the aid of homogeneous vanadium and metallocene catalysts were compared by DSC and time-resolved simultaneous SAXS-WAXS-DSC at scanning rates of 10 and 20°C min?1 using synchrotron radiation. An EP copolymer with a density of 896 kg m?3 (about 89 mol % ethylene) after compression moulding gave orthorhombic WAXS reflections. The crystallinity as a function of temperature [w c (T)] calculated from these reflections using the two-phase model was in good agreement withw c (T) calculated fromc p measurements using DSC. Thec p measurements also enabled calculation of the baselinec p and the excessc p. The SAXS measurements revealed a strong change in the long period in cooling and in heating. The SAXS invariant as a function of temperature showed a maximum in both cooling and heating, which could be explained from the opposing influences of the crystallinity and the electron density difference between the two phases. Two EO copolymers with densities of about 871 kg m?3 (about 87 mol% ethylene) no longer showed any clear WAXS reflections, although DSC and SAXS measurements showed that these copolymers did crystallize. The similarity between the results led to the conclusion that the copolymers, though based on different catalyst systems — vanadium and metallocene — did not have strongly different sets of propagation probabilities of chain growth during polymerization. On the basis of a Monte Carlo simulation model of crystallization and morphology, based on detailed knowledge of the microchain structure, the difference between WAXS on the one hand and DSC and SAXS on the other could be explained as being due to loosely packed crystallized ethylene sequences in clusters. These do cause the density and the electron density of the cluster to increase (which is measurable by SAXS) and the enthalpy to decrease (which is measurable by DSC) but the clusters are too small and/or too imperfect to give constructive interference in the case of WAXS. Of an EP copolymer with an even lower ethylene content (about 69 mol %), the crystallization and melting processes could still be readily measured by DSC and SAXS, which proves that these techniques are eminently suitable for investigating the crystallization and melting behaviour of the copolymers studied.  相似文献   

13.
The multiple melting behavior of poly(butylene succinate) (PBSu) was studied with differential scanning calorimetry (DSC). Three different PBSu resins, with molecular weights of 1.1 × 105, 1.8 × 105, and 2.5 × 105, were cooled from the melt (150 °C) at various cooling rates (CRs) ranging from 0.2 to 50 K min?1. The peak crystallization temperature (Tc) of the DSC curve in the cooling process decreased almost linearly with the logarithm of the CR. DSC melting curves for the melt‐crystallized samples were obtained at 10 K min?1. Double endothermic peaks, a high‐temperature peak H and a low‐temperature peak L, and an exothermic peak located between them appeared. Peak L decreased with increasing CR, whereas peak H increased. An endothermic shoulder peak appeared at the lower temperature of peak H. The CR dependence of the peak melting temperatures [Tm(L) and Tm(H)], recrystallization temperature (Tre), and heat of fusion (ΔH) was obtained. Their fitting curves were obtained as functions of log(CR). Tm(L), Tre, and ΔH decreased almost linearly with log(CR), whereas Tm(H) was almost constant. Peak H decreased with the molecular weight, whereas peak L increased. It was suggested that the rate of the recrystallization decreased with the molecular weight. Tm(L), Tm(H), Tre, and Tc for the lowest molecular weight sample were lower than those for the others. In contrast, ΔH for the highest molecular weight sample was lower than that for the others. If the molecular weight dependence of the melting temperature for PBSu is similar to that for polyethylene, the results for the molecular weight dependence of PBSu can be explained. © 2002 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 40: 2411–2420, 2002  相似文献   

14.
Non-isothermal crystallization of isotactic poly(4-methyl-pentene-1) (P4MP1) is studied by differential scanning calorimeter (DSC), and kinetic parameters such as the Avrami exponent and the kinetic crystallization rate (Z c) are determined. From the cooling and melting curves of P4MP1 at different cooling rates, the crystalline enthalpy increases with the increasing cooling rate, but the degree of crystalline by DSC measurement shows not much variation. Degree of crystalline of P4MP1 calculated by wide angle X-ray diffraction pattern shows the same tendency with crystalline enthalpy, indicating that re-crystallization occurs when samples heated above the second glass transition temperature of P4MP1. By Jeziorny analysis, n 1 value suggests that mainly spherulites’ growth at 2.5 K min−1 transforms into a mixture mode of three-dimensional and two-dimensional space extensions with further increasing cooling rate. In the secondary crystallization process, n 2 values indicate that the secondary crystallization is mainly the two-dimensional extension of the lamellar crystals formed during the primary crystallization process. The rates of the crystallization, Z c and t 1/2 both increase obviously with the increase of cooling rate, especially at the primary crystallization stage. By Mo’s method, higher cooling rate should be required in order to obtain a higher degree of crystallinity at unit crystallization time.  相似文献   

15.
We report on a detailed analysis of the thermal transition hysteresis loop and the relaxation curves of the recently investigated compound [FeII(L)2][ClO4]2 · C7H8 [V. Mishra et al., Inorg. Chem. (2008)]. We developed a variant of a previous kinetic cooperative model, modified for including short-range and long-range Gaussian distributions of energy parameters. The resulting set of coupled master equations was solved numerically for the dynamic and the quasi-static case. We have elucidated the various correlations between the distribution widths and the long-range interaction parameter, which basically have opposite effects. It is shown that presence of a steep hysteresis loop may hide a sizable short-range distribution, while the stretched character of the relaxation curves may also hide the presence of strong interactions. We also show that the coherent analysis of the set of relaxation curves at different temperatures does not allow to disentangle properly all the relevant parameters. We discuss these data with respect to the other photomagnetic experiments, T(LIESST) curve and light-induced thermal hysteresis (LITH) loop, and finally we try to address the question of the real nature of the cooperative relaxation process.  相似文献   

16.
1-Ethyl-3-methylimidazolium acetate was studied by NMR relaxation. The temperature dependences of the spin-lattice relaxation rates (1/T 1) for 1H and 13C were obtained. The curves with maxima were observed for the majority of the temperature dependences 1/T 1, which provided a reliable temperature dependence of the correlation times (τc). In the low-temperature range, the proton relaxation rates tend to an asymptotic value, which is related, most likely, to spin diffusion manifested in the studied samples. The values of correlation times τc calculated for 1H and 13C of the same functional group almost coincide at high temperatures, which confirms that the used approach is adequate for the determination of characteristic times of rotational reorientation of counterions in the studied ionic liquid.  相似文献   

17.
The thermal effect accompanying the transition of Cu2–xSe into a superionic conduction state was studied by non-isothermal measurements, at different heating and cooling rates (β=1, 2.5, 5, 10 and 20°C min–1). During heating the peak temperature (Tp) remains almost stable for all values of β, (136.8±0.4°C for Cu2Se and 133.0±0.3°C for Cu1.99Se). A gradual shift of the initiation of the transformation towards lower temperatures is observed, as the heating rate increases. During cooling there is a significant shift in the position of the peak maximum (Tp) towards lower temperatures with the increase of the cooling rate. A small hysteresis is observed, which increases with the increase of the cooling rate, β. The mean value of transformation enthalpy was found to be 30.3±0.8 J g–1 for Cu2Se and 28.9±0.9 J g–1 for Cu1.99Se. The transformation can be described kinetically by the model f(ǯ)=(1–ǯ)n(1+kcatX), with activation energy E=175 kJ mol–1, exponent value n equal to 0.2, logA=20 and log(kcat)= 0.5.  相似文献   

18.
The dielectric spectra of polypropyleneglycols H-(C3H6O) N p -OH (PPGs), where N p = 1, 2, 3, 7, 12, 17, 20, 34, 69, were analyzed in terms of the Dissado Hill (DH) cluster model above the vitrification temperatures. In PPGs, the structural clusters are associates formed by intra- and intermolecular hydrogen bonds. The activation processes of cleavage and formation of intermolecular hydrogen bonds in clusters, when the total number of intermolecular hydrogen bonds changes, are characterized by the parameter n DH. The fluctuation processes of simultaneous exchange of molecules between adjacent clusters correspond to redistributions of intermolecular hydrogen bonds between clusters, when only the position but not the total number of intermolecular hydrogen bonds changes, and are characterized by the parameter m DH. The relaxation time τDH at 303 K and 423 K and the parameters n DH and m DH of the dielectric spectra were calculated. The activation energies of relaxation in the range 210–323 K were determined. The mean statistic squares of the dipole moments of clusters 〈μc2〉 and di-PG, PPG-425 (N p = 7), and PPG-2025 (N p = 34) molecules 〈μm2〉 at 303 K and 423 K were calculated. The number of the units of the oxypropylene chains involved in relaxation was determined. The dependence of the parameters of the DH model, relaxation energies, 〈μc2〉 and 〈μm2〉 on N p were studied.  相似文献   

19.
To continue dynamicc p measurements in the range of smallest temperature rates and non-linear thermal relaxation investigations into the linear range, simultaneousc p and thermal relaxation measurements were carried out in an adiabatic vacuum calorimeter, using the pulse heating method. The rate-dependentc p behaviour, known from dynamic measurements, does not continue at small temperature rates. This is confirmed by the relaxation process which is observed. The results suggest an extended interpretation of the glass transition in atactic polymethylmethacrylate (PMMA).  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