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 546 毫秒
1.
边六交  a  王晓华b  杨晓燕a  刘莉a 《中国化学》2008,26(7):1276-1284
通过用前沿分析测定热力学参数的方法,研究了枯草杆菌α-淀粉酶在几种色谱介质上的热变性行为。实验结果表明,在RP-C18反相介质、Zn2+螯合的Chelating Sepharose Fast-Flow亲和介质和WCX-1阳离子交换介质上,当温度分别低于或超过30℃时,α-淀粉酶分子分别以一种稳定的构象存在;而在PEG-400 和修饰的PEG-400疏水色谱介质上,当温度分别低于40℃和30℃时,α-淀粉酶分子分别以一种稳定的构象存在,但当温度分别高于40℃和30℃时,α-淀粉酶分子的构象会发生剧烈的变化。同时,通过测定α-淀粉酶分子在自由溶液以及在PEG-400 和修饰的PEG-400疏水色谱介质上的热失活曲线,可以得出结论:在液相色谱过程中,色谱介质会诱导α-淀粉酶分子构象的变化,并促进它们的热变性;而在疏水色谱中,色谱介质的疏水性越高,α-淀粉酶分子在其上的构象变化温度越低。  相似文献   

2.
在15~85℃宽温度范围,研究了蛋白质在固定Zn2 金属螯合色谱系统中的热行为和变性热力学。实验结果表明,蛋白质在色谱过程都有一个固定的热转变温度:核糖核酸酶(RNase)、α-胰凝乳蛋白酶原A(α-Chy)的热转变温度约为55℃,细胞色素C(Cyt-C)和溶菌酶(Lys)约为65℃;,热转变温度的出现标志蛋白质构象发生变化;利用Van′tHoff作图测定了蛋白质在色谱系统热变性时的标准焓变ΔH°和标准熵变ΔS°,提出用标准熵变ΔS°和自由能变ΔG°判断蛋白质构象变化;利用ΔH°-ΔS°的线性关系估算了蛋白质热变性时的补偿温度,鉴定了蛋白质各变体在金属螯合色谱中保留机理的同一性,RNase、Cyt-C、Lys和α-Chy的补偿温度分别为55℃、65.8℃、65.2℃和54.8℃;根据蛋白质热变性时的补偿温度和构象变化熵变Δ(ΔS°)的大小,讨论了蛋白质在阳离子交换色谱和固定Zn的金属螯合色谱体系中的热稳定性。实验证明,在IDA裸柱引入Zn2 后蛋白质在色谱系统中的热稳定性减小,平均补偿温度从65.3℃降低到59.7℃,而构象变化熵变的绝对值大幅度升高。  相似文献   

3.
研究了0~50℃内五种蛋白质在疏水色谱(HIC)中的保留行为。蛋白分子在HIC中的保留行为遵循Van'tHoff方程,并求得蛋白质分子在HIC中保留的热力学参数(△H°,△S°,△C°p和△G°)。发现蛋白质的保留由熵控制,焓、熵和热容量三个热力学参数都与绝对温度倒数间存在线性关系,并求得热容量为零的收敛温度。  相似文献   

4.
液相色谱过量热力学函数变化规律性探讨   总被引:1,自引:0,他引:1  
侯镜德  徐秀珠  王国庆 《化学学报》1988,46(10):961-966
本文从溶液热力学和色谱理论推导出过量热力学函数与色谱参数间的关系, 测定了苯的同系物在液-液分配过程中, 过量焓变等函数与流动相组成、样品碳数的线性关系,并求得经验方程式. 同时研究了不同样品焓变和熵变的线性补偿规律. 并试图从溶液理论进行计算.  相似文献   

5.
依据液相色谱中溶质计量置换保留模型, 对溶质在反相液相色谱(RPLC)保留过程及其吸附、解吸附过程中的焓熵补偿进行了研究, 证实了在RPLC中焓熵补偿确实存在。从焓熵补偿的定义出发, 从理论上证明了溶质在保留过程中的焓熵补偿温度本质上为溶质保留值的收敛温度, 其数值为Z对1/T线性作图的斜率与截距之比。与惯常计算焓熵补偿温度的方法相比, 本文的方法所得补偿温度更为合理且不受流动相中强溶剂浓度变化的影响。  相似文献   

6.
白泉  张瑞燕  耿信笃 《化学学报》1997,55(10):1025-1029
依据液相色谱中溶质计量置换保留模型, 对溶质在反相液相色谱(RPLC)保留过程及其吸附、解吸附过程中的焓熵补偿进行了研究, 证实了在RPLC中焓熵补偿确实存在。从焓熵补偿的定义出发, 从理论上证明了溶质在保留过程中的焓熵补偿温度本质上为溶质保留值的收敛温度, 其数值为Z对1/T线性作图的斜率与截距之比。与惯常计算焓熵补偿温度的方法相比, 本文的方法所得补偿温度更为合理且不受流动相中强溶剂浓度变化的影响。  相似文献   

7.
用气液色谱法,在不同的温度下,测定了以α-甲基萘作为固定液时,路易斯碱脂肪胺和醇的溶解平衡常数K_R~0,以及它们与作为路易斯酸的过渡金属配合物双-(4,6-二异丙基水杨酸)合铜(Ⅱ)相互作用的表观分配常数K_R和加合反应的平衡常数K_1。再根据热力学公式: 进一步求出了加合反应的焓变ΔH和熵变ΔS。  相似文献   

8.
研究21-80℃温度范围内一些蛋白质和小分子在疏水相互作用色谱中的热行为。利用Van't Hoff作图(lnk'-1/T)测定蛋白质分子的热力学参数(ΔH°, ΔS°和ΔG°), 根据标准熵变(ΔS°)和标准自由能变(ΔG°)判断蛋白质在色谱过程中的构象变化, 通过ΔH°-ΔS°的线性关系估计蛋白质变性时的"补偿温度"(β), 鉴定蛋白质在疏水相互作用色谱中保留机理的同一性。  相似文献   

9.
液-固吸附体系中计量置换吸附模型的热力学研究   总被引:13,自引:0,他引:13  
研究了液-固吸附体系中溶质计量置换吸附过程中的自由能变,推导出了新的表示溶质种类、溶质浓度和溶剂对Gibbs自由能变贡献的数学表达式,并阐明了该表达式中各种参数的物理意义.将用通常方法测定的自由能变分为与溶质吸附和溶剂解吸附有关的两个独立的自由能变,推导出了液-固吸附体系中溶质计量置换吸附模型的两个线性参数βa和q/Z对绝对温度倒数的线性关系,又将吸附过程中溶质的焓变和熵变分成两个独立的分量.用文献中的实验数据,对本文推导出的公式进行了检验,理论结果与实验结果一致.  相似文献   

10.
王文英 《化学教育》2016,37(15):71-72
运用化学反应的焓变与熵变,从化学热力学的角度,计算说明了如果空气中二氧化碳的含量按文献值即体积浓度为0.03%上下各浮动20%,求得碳酸钠分解温度为1227.69℃和1203.74℃,用酒精灯加热不能达到;碳酸氢钠分解温度为54.56℃和50.9℃,用酒精灯加热很容易达到;故用酒精灯加热只能使碳酸氢钠分解产生二氧化碳,而碳酸钠则不能。  相似文献   

11.
The thermodenaturation behavior of Bacillus subtilis α‐amylase on some chromatographic media was studied by determining their adsorption parameters with frontal analysis. The experimental results show that on a RP‐C18 reversed‐phase medium, a Chelating Sepharose Fast‐Flow chelated by Zn2+ affinity medium and a WCX‐1 cation‐exchange medium, a stable conformation of α‐amylase molecule separately exists below or over 30 °C; while on a PEG‐400 hydrophobic medium and a modified PEG‐400 medium, a stable conformation of α‐amylase molecule separately exists below 40 and 30 °C, and when the experimental temperatures are separately over 40 and 30 °C, a drastically conformational change of α‐amylase molecules can continuously take place. And by combining the intrinsic fluorescence emission spectrum and thermal inactivation profile of α‐amylase in free solution and on the PEG‐400 and modified PEG‐400 hydrophobic media, it can be concluded that in liquid chromatographic procedure, chromatographic media can induce the conformational change of α‐amylase molecules and promote their thermodenaturation; and in hydrophobic interaction chromatography, the higher the hydrophobicity of chromatographic medium, the lower the conformational change temperature of α‐amylase molecules on the chromatographic medium.  相似文献   

12.
A doubly hydrophilic triblock copolymer of poly(N,N‐dimethylamino‐2‐ethyl methacrylate)‐b‐Poly(ethylene glycol)‐b‐poly(N,N‐dimethylamino‐2‐ethylmethacrylate) (PDMAEMA‐b‐PEG‐b‐PDMAEMA) with well‐defined structure and narrow molecular weight distribution (Mw/Mn = 1.21) was synthesized in aqueous medium via atom transfer radical polymerization (ATRP) of N,N‐dimethylamino‐2‐ethylmethacrylate (DMAEMA) initiated by the PEG macroinitiator. The macroinitiator and triblock copolymer were characterized with 1H NMR and gel permeation chromatography (GPC). Fluorescence spectroscopy, dynamic light scattering (DSL), transmittance measurement, and rheological characterization were applied to investigate pH‐ and temperature‐induced micellization in the dilute solution of 1 mg/mL when pH > 13 and gelation in the concentrated solution of 25 wt % at pH = 14 and temperatures beyond 80 °C. The unimer of Rh = 3.7 ± 0.8 nm coexisted with micelle of Rh = 45.6 ± 6.5 nm at pH 14. Phase separation occurred in dilute aqueous solution of the triblock copolymer of 1 mg/mL at about 50 °C. Large aggregates with Rh = 300–450 nm were formed after phase separation, which became even larger as Rh = 750–1000 nm with increasing temperature. The gelation temperature determined by rheology measurement was about 80 °C at pH 14 for the 25 wt % aqueous solution of the triblock copolymer. © 2008 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 46: 5869–5878, 2008  相似文献   

13.
The introduction of RbF into the Mg(NH2)2–2 LiH system significantly decreased its (de‐)hydrogenation temperatures and enhanced its hydrogen‐storage kinetics. The Mg(NH2)2–2 LiH–0.08 RbF composite exhibits the optimal hydrogen‐storage properties as it could reversibly store approximately 4.76 wt % hydrogen through a two‐stage reaction with the onset temperatures of 80 °C for dehydrogenation and 55 °C for hydrogenation. At 130 °C, approximately 70 % of hydrogen was rapidly released from the 0.08 RbF‐doped sample within 180 min, and the fully dehydrogenated sample could absorb approximately 4.8 wt % of hydrogen at 120 °C. Structural analyses revealed that RbF reacted readily with LiH to convert to RbH and LiF owing to the favorable thermodynamics during ball‐milling. The newly generated RbH participated in the following dehydrogenation reaction, consequently resulting in a decrease in the reaction enthalpy change and activation energy.  相似文献   

14.
For the preparation of well‐defined H2O‐soluble C60 polymers, several C60‐PEG conjugates were prepared from a C60 biscarboxylic acid derivative and monodisperse NH2‐PEGs (NH2‐EGn, = 4 – 36) via amide conjugation. When the relatively long PEGs (EGn,  12) were employed, the C60‐PEG conjugates became completely H2O‐soluble by forming micelle‐like structure shown by the data of surface tension, DLS, and cryo‐TEM. Interestingly, these H2O‐soluble C60‐PEG conjugates (C60(EGn)2, = 12 – 36) showed reversible thermoresponse to form larger aggregates (ca. 1 μm by DLS) at higher temperatures. The temperature for the aggregation was related to the lengths of PEGs attached to C60; 29 °C (C60(EGn)2, = 12), 51 °C (= 20), and 72 °C (= 36). This thermoresponse was speculated to occur by dehydration of well‐organized PEG chains in the micelle‐type structure of monodisperse C60‐PEG caused by gauche‐to‐anti conformational change of PEG anchors. This thermoresponse of well‐defined amphiphilic C60‐PEG conjugates indicates potential applications in areas such as temperature sensors and thermoresponsive materials.  相似文献   

15.
Ring‐opening metathesis polymerization of 4‐phenylcyclopentene is investigated for the first time under various conditions. Thermodynamic analysis reveals a polymerization enthalpy and entropy sufficient for high molar mass and conversions at lower temperatures. In one example, neat polymerization using Hoveyda–Grubbs second generation catalyst at −15 °C yields 81% conversion to poly(4‐phenylcyclopentene) (P4PCP) with a number average molar mass of 151 kg mol−1 and dispersity of 1.77. Quantitative homogeneous hydrogenation of P4PCP results in a precision ethylene‐styrene copolymer (H2‐P4PCP) with a phenyl branch at every fifth carbon along the backbone. This equates to a perfectly alternating trimethylene‐styrene sequence with 71.2% w/w styrene content that is inaccessible through molecular catalyst copolymerization strategies. Differential scanning calorimetry confirms P4PCP and H2‐P4PCP are amorphous materials with similar glass transition temperatures (Tg) of 17 ± 2 °C. Both materials present well‐defined styrenic analogs for application in specialty materials or composites where lower softening temperatures may be desired.

  相似文献   


16.
17.
Anionic polymerizations of three 1,3‐butadiene derivatives containing different N,N‐dialkyl amide functions, N,N‐diisopropylamide (DiPA), piperidineamide (PiA), and cis‐2,6‐dimethylpiperidineamide (DMPA) were performed under various conditions, and their polymerization behavior was compared with that of N,N‐diethylamide analogue (DEA), which was previously reported. When polymerization of DiPA was performed at ?78 °C with potassium counter ion, only trace amounts of oligomers were formed, whereas polymers with a narrow molecular weight distribution were obtained in moderate yield when DiPA was polymerized at 0 °C in the presence of LiCl. Decrease in molecular weight and broadening of molecular weight distribution were observed when polymerization was performed at a higher temperature of 20 °C, presumably because of the effect of ceiling temperature. In the case of DMPA, no polymer was formed at 0 °C and polymers with relatively broad molecular weight distributions (Mw/Mn = 1.2) were obtained at 20 °C. The polymerization rate of PiA was much faster than that of the other monomers, and poly(PiA) was obtained in high yield even at ?78 °C in 24 h. The microstructure of the resulting polymers were exclusively 1,4‐ for poly(DMPA), whereas 20–30% of the 1,2‐structure was contained in poly(DiPA) and poly(PiA). © 2010 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 48: 3714–3721, 2010  相似文献   

18.
Cellulose ( 1 ) was converted for the first time to 6‐phenyl‐6‐deoxy‐2,3‐di‐O‐methylcellulose ( 6 ) in 33% overall yield. Intermediates in the five‐step conversion of 1 to­ 6 were: 6‐O‐tritylcellulose ( 2 ), 6‐O‐trityl‐2,3‐di‐O‐methylcellulose ( 3 ), 2,3‐di‐O‐methylcellulose ( 4 ); and 6‐bromo‐6‐deoxy‐2,3‐di‐O‐methylcellulose ( 5 ). Elemental and quantitative carbon‐13 analyses were concurrently used to verify and confirm the degrees of substitution in each new polymer. Gel permeation chromotography (GPC) data were generated to monitor the changes in molecular weight (DPw) as the synthesis progressed, and the compound average decrease in cellulose DPw was ~ 27%. Differential scanning calorimetry (DSC) and thermogravimetric analysis (TGA) were used to characterize the decomposition of all polymers. The degradation temperatures ( °C) and percent char at 500 °C of cellulose derivatives 2 to 6 were 308.6 and 6.3%, 227.6 °C and 9.7%, 273.9 °C and 30.2%, 200.4 °C and 25.6%, and 207.2 °C and 27.0%, respectively. The glass transition temperature (Tg) of­6‐O‐tritylcellulose by dynamic mechanical analysis (DMA) occurred at 126.7 °C and the modulus (E′, Pa) dropped 8.9 fold in the transition from ?150 °C to + 180 °C (6.6 × 109 to 7.4 × 108 Pa). Modulus at 20 °C was 3.26 × 109 Pa. Complete proton and carbon‐13 chemical shift assignments of the repeating unit of the title polymer were made by a combination of the HMQC and COSY NMR methods. Ultimate non‐destructive proof of carbon–carbon bond formation at C6 of the anhydroglucose moiety was established by generating correlations between resonances of CH26 (anhydroglucose) and C1′, H2′, and H6′ of the attached aryl ring using the heteronuclear multiple‐bond correlation (HMBC) method. In this study, we achieved three major objectives: (a) new methodologies for the chemical modification of cellulose were developed; (b) new cellulose derivatives were designed, prepared and characterized; (c) unequivocal structural proof for carbon–carbon bond formation with cellulose was derived non‐destructively by use of one‐ and two‐dimensional NMR methods. Copyright © 2002 John Wiley & Sons, Ltd.  相似文献   

19.
The enantiomeric pairs of cis and trans stereoisomers of cyclic β‐aminohydroxamic acids and their related cis and trans cyclic β‐amino acids containing two chiral centers were directly separated on four structurally related chiral stationary phases derived from quinine and quinidine modified with (R,R)‐ and (S,S)‐aminocyclohexanesulfonic acids. Applying these zwitterionic ion‐exchangers as chiral selectors, the effects of the composition of the bulk solvent, the acid and base additives, the structures of the analytes, and temperature on the enantioresolution were investigated. To study the effects of temperature and obtain thermodynamic parameters, experiments were carried out at constant mobile phase compositions in the temperature range 5–50°C. The differences in the changes in standard enthalpy Δ(ΔH°), entropy Δ(ΔS°), and free energy Δ(ΔG°) were calculated from the linear van't Hoff plots derived from the ln α versus 1/T curves in the studied temperature range. Results thus obtained indicated enthalpy‐driven separations in all cases. The sequence of elution of the enantiomers was determined and found to be reversed when ZWIX(–)™ was changed to ZWIX(+)™ or ZWIX(–A) to ZWIX(+A).  相似文献   

20.
A series of comb polymers consisting of a methacrylate backbone and poly(2‐ethyl‐2‐oxazoline) (PEtOx) side chains was synthesized by a combination of cationic ring‐opening polymerization and reversible addition–fragmentation chain transfer polymerization. Small‐angle neutron scattering (SANS) studies revealed a transition from an ellipsoidal to a cylindrical conformation in D2O around a backbone degree of polymerization of 30. Comb‐shaped PEtOx has lowered Tg values but a similar elution behavior in liquid chromatography under critical conditions in comparison to its linear analog was observed. The lower critical solution temperature behavior of the polymers was investigated by turbidimetry, dynamic light scattering, transmission electron microscopy, and SANS revealing decreasing Tcp in aqueous solution with increasing molar mass, the presence of very few aggregated structures below Tcp, a contraction of the macromolecules at temperatures 5 °C above Tcp but no severe conformational change of the cylindrical structure. In addition, the phase diagram including cloud point and coexistence curve was developed showing an LCST of 75 °C of the binary mixture poly[oligo(2‐ethyl‐2‐oxazoline)methacrylate]/water. © 2012 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2013  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