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 31 毫秒
1.
The conformational energy contribution (ΔUconf) to the heat of solution in polymer-solvent systems is presented and discussed in connection with chain conformational properties. In particular, ΔUconf has been discussed in terms of various possible mechanisms of coil deformation.  相似文献   

2.
Heats of mixing and excess volumes at infinite dilution have been obtained at 25°C. for polydimethylsiloxane or its lower oligomers in various solvents by using a twin conduction microcalorimeter and from the pycnometric specific volumes. From those values, excess energies ΔEMv at constant volume have been determined. The prediction on intramolecular conformation contributions to the heat of solution as proposed by Bianchi has been evaluated by the values of ΔEMv. The heat of solution in the polymer–solvent systems was interpreted by the expression for ΔEMv derived from the Van Laar-Hildebrand work on simple liquid mixtures with the solubility parameters of polymers obtained from indirect measurements. The values of conformational intramolecular energy change calculated from dilute solution properties were difficult to rationalize with our results. Our present results suggest that systems in a nonideal state can not be distinguished for certain from those in the ideal state. This conclusion based on apparent values does not deny the possible effect of the conformational energy change.  相似文献   

3.
The surface characterization of 2‐(dimethylamino)ethylmethacrylate (DMA) and 2‐(N‐morpholino)ethylmethacrylate (MEMA) homopolymers and DMA–MEMA diblock copolymer was studied using inverse‐gas chromatography (IGC). The analyzed surface properties of (co)polymers were the dispersive component of the surface energy ( ) and the acid–base characters of (co)polymer surfaces. The specific free energy (ΔGsp), enthalpy (ΔHsp), and entropy (ΔSsp) of adsorption of polar probes on (co)polymers were calculated. The values of ΔHsp were correlated with both the donor and the modified acceptor numbers (AN) of the probes to quantify the acidic KA and the basic KD parameters of (co)polymer surfaces. The values obtained for the KA and KD parameters indicated basic characters for the surface of (co)polymers. The dispersive component values of the surface energy and the acid–base surface parameters of the DMA–MEMA diblock copolymer surface were found to be between those homopolymers as expected. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

4.
Abstract

Twinned dimeric mesogens having a rigid-flexible-rigid molecular structure have been shown to be appropriate models for some properties of regularly alternating (rigid-flexible)n main chain liquid crystalline polymers (lcps). A family of tetrameric monodisperse liquid crystalline model compounds chemically related to known main chain liquid crystalline polymers of the 4-alkoxyphenyl 4′-alkoxy-benzoate type has been synthesized. The tetramers are nematogenic. Alternations in thermodynamic parameters (ΔH, ΔS) for the N-I transition as a function of spacer chain length indicate conformational behaviour of the internal spacers dominates mesophase properties.  相似文献   

5.
M. Lotfi  R.M.G. Roberts 《Tetrahedron》1979,35(18):2131-2136
The rates of addition of tetracyanoethylene to a number of 9- and 9,10-substituted anthracenes have been measured spectrophotometrically in solvent CCl4. Substituent effects correlated well using the extended form of the Hammett equation. The importance of steric effects on the reaction was assessed by a systematic variation of the components of the data used in the above correlation.Activation parameters (ΔGexp, etc.) and the corresponding overall thermodynamic parameters for adduct formation (ΔGad°, etc.) were evaluated. ΔGexp was found to be linearly related to ΔGc°, the free energy of formation of the intermediate complex which confirms the role of the latter as a true reaction intermediate. From correlations between ΔGexp and ΔGad°, an “early” transition state is suggested. The above thermodynamic and activation data enable detailed reaction profiles to be drawn.  相似文献   

6.
The difference between the expectation values of the total electronic kinetic energy operator (ΔEK), and the operators accounting for the Coulombic interactions between the electrons and nuclei (ΔVen), between all pairs of electrons (ΔVee), and between all pairs of nuclei (ΔVnn) for the product and reactant species in a wide variety of hydrocarbon reactions are calculated using single determinant basis set data reported in the literature. Following Allen, their contributions to ΔET, the difference between the corresponding total molecular energies and thus the reaction heat, are grouped together as a repulsion energy term, ΔErep = ΔEK + ΔVee + ΔVnn, and an attraction energy term ΔEattr = ΔVen. For all but 2 of the 71 individual reactions considered in this paper, the experimental reaction heat at 0°K corrected for zero-point energy contributions, (ΔH)zpe, is the result of near compensation between far larger ΔErep and ΔEattr terms, in sharp contrast to the much smaller ΔErep and ΔEattr terms which are characteristic of many molecular rotation processes. By matching the sign of (ΔH)zpe with that of ΔErep or ΔEattr, as the case may be, the reactions are classified as attractive-dominant or repulsive-dominant (46 in the former class and 23 in the latter), a property which is independent of the direction in which the reaction is written. The sign and magnitude of ΔVee, ΔVnn, and ΔVen and reaction category are discussed in relation to the various kinds of structural change involved in going from reactants to products. For the vast majority of reactions, the numerical relationship ΔVee ≈ ΔVnn has been found to hold to within a few percent.  相似文献   

7.
The heat of fusion (ΔHu) of isotactic poly(methyl methacrylate) (PMMA) was determined from the combined linear regression analysis of x-ray diffraction data and differential scanning calorimetry. By means of two sample preparation techniques to reduce crystallinity, a thermal pretreatment in which a partial melt and quench cycle was utilized, and a mechanical mixture in which a semicrystalline and amorphous material were blended, a value for ΔHu = 1200 ± 80 cal/mole (100 g) was found. The present result represents the first measurement of ΔHu for an acrylic polymer and yields G(-units) = 12 from cryoscopic measurements for the chemical changes caused by high energy irradiation.  相似文献   

8.
2HNMR measurements were performed on main-chain dimer and polymer liquid crystals (LC) having oxyethylene (OE) spacers -(OCH2CH2)xO-(x=2, 3). The orientational as well as conformational characteristics of these molecules have been investigated in bulk and in a nematic solution. The OE spacer was found to take spatial arrangements characteristics of the nematic phase. The nematic conformation of the spacer remains nearly invariant over a wide range of temperature and concentration. In these analyses, the ratio of the deuterium quadrupolar splittings ΔννRν: spacer and ΔνR: mesogenic unit) provided an important information regarding the spatial configuration of molecules in the LC state. The results obtained in this study are consistent with our previous conclusion drawn on a series of main-chain LC oligomers and polymers comprising n-alkane spacers -O(CH2)nO- (n=9, 10).  相似文献   

9.
The anionic polymerization of three monomers, 2-isopropenyl-4,5-dimethyloxazole(I), 2-isopropenylthiazole(II), and 2-isopropenylpyridine(III), was studied in THF. These monomers produced red-colored living polymers on addition of sodium naphthalene or living α-methylstyrene tetramer as an initiator. It was observed that a considerable amount of monomer remained in the respective living polymer–monomer system, indicating that an equilibrium between the polymer and the monomer existed as in the case of α-methylstyrene. At lower temperatures, the conversion of the monomer to the polymer increased. The equilibrium monomer concentrations [Me] were determined at different temperatures, and the heats (ΔH) and the entropies (ΔS°) of polymerization were obtained by plotting In(1/[Me]) against 1/T as ΔH = ?9.4, ?6.8, and ?6.2 kcal/mole, ΔS°S = ?22.9, ?16.5, and ?16.6, eu for I, II, and III, respectively.  相似文献   

10.
11.
For BSA and β-lactoglobulin adsorption to hydrophobic interaction chromatography (HIC) stationary phases leads to conformational changes. In order to study the enthalpy (ΔHads), entropy (ΔSads), free energy (ΔGads) and heat capacity (Δcp,ads) changes associated with adsorption we evaluated chromatographic data by the non-linear van’t Hoff model. Additionally, we performed isothermal titration calorimetry (ITC) experiments. van’t Hoff analysis revealed that a temperature raise from 278 to 308 K increasingly favoured adsorption seen by a decrease of ΔGads from −12.9 to −20.5 kJ/mol for BSA and from −6.6 to −13.2 kJ/mol for β-lactoglobulin. Δcp,ads values were positive at 1.2 m (NH4)2SO4 and negative at 0.7 m (NH4)2SO4. Positive Δcp,ads values imply hydration of apolar groups and protein unfolding. These results further corroborate conformational changes upon adsorption and their dependence on mobile phase (NH4)2SO4 concentration. ITC measurements showed that ΔHads is dependent on surface coverage already at very low loadings. Discrepancies between ΔHads determined by van’t Hoff analysis and ITC were observed. We explain this with protein conformational changes upon adsorption which are not accounted for by van’t Hoff analysis.  相似文献   

12.
Solubility data for poly(3‐hexylthiophene) (P3HT) in 29 pure solvents are presented and discussed in detail. Functional solubility parameter (FSP) and convex solubility parameter (CSP) computations are performed and the CSP and FSP results are compared to previously reported Hansen solubility parameters (HSPs) and to the parameters calculated using additive functional group contribution methods. The empirical data reveals experimental solubility parameters with substantial polar (δP) and hydrogen‐bonding (δH) components, which are not intrinsic to the structure of the P3HT polymer. Despite these apparent irregularities, it is shown that the predictor method based on the solubility function, f, does provide a reliable way to quantitatively evaluate the solubility of P3HT in other solvents in terms of a given set of empirical solubility data. The solubility behavior is further investigated using linear solvation energy relationship (LSER) modeling and COSMO‐RS computations of the activity coefficients of P3HT. The LSER model reveals that (1) the cavity term, δT, is the dominant factor governing the solubility behavior of P3HT and (2) the solvent characteristics that dictate the structural order (crystallinity) of P3HT aggregates do not similarly influence the overall solubility behavior of the polymer. © 2017 Wiley Periodicals, Inc. J. Polym. Sci., Part B: Polym. Phys. 2017 , 55, 1075–1087  相似文献   

13.
The solvent dependence of thermodynamic parameters of conformational equilibria in trans-1,2-dichlorocyclohexane and trans-1,2-bromochlorocyclohexane was investigated by infrared absorption spectra. The results obtained show the existence of a compensation effect in the thermodynamics of conformational equilibria: the enthalpy (ΔH0) and entropy (ΔS0) differences change in the same direction when going from one solvent to another. A semi-quantitative estimation of the effect is given on the basis of the equations of statistical thermodynamics. It is shown that the temperature dependence of the ΔS0 value must be taken into account when determining the enthalpy difference of the conformers. This yields the equality of the true and observed ΔH0 values.  相似文献   

14.
The enthalpies of solution in water for five new light rare earth ternary complexes RE(Gly)4Im(ClO4)3 2H2O (RE = La, Pr, Nd, Sm, Eu; Gly‐glycine; Im‐imidazole) were measured by means of a Calvet microcalorimeter. The empirical formula of enthalpy of solution (ΔsolH), relative apparent molar enthalpy (πLi), relative partial molar enthalpy (Li) and enthalpy of dilution (ΔdllH1,2) were drawn up by the data of enthalpies of solution of these complexes. From three plots of the values of standard enthalpy of solution Δsol H?, πLi, Li) versus the values of ionic radius (r) of the light rare earth elements, the grouping effect of lanthanide was observed, showing that the coordination bond between rare earth ion and ligand possesses a certain extent of the property of a covalent bond. The standard enthalpies of solution in water of similar complexes, Ce(Gly)4Im(ClO4)3.2H2O were estimated according to the plot of ΔsolH?, versus r.  相似文献   

15.
The volumes of mixing of hexadecane and each of the isomers of hexane have been measured for the equimolar mixtures at 20°C. The results have been used together with previously measured values of ΔH to obtain ΔUv. A very good correlation is found between the energy of mixing and the properties of the pure alkanes.  相似文献   

16.
Dissociation constants (pKa) of trazodone hydrochloride (TZD⋅HCl) in EtOH/H2O media containing 0, 10, 20, 30, 40, 50, 60, 70, and 80% (v/v) EtOH at 288.15, 298.15, 308.15, and 318.15 K were determined by potentiometric techniques. At any temperature, pKa decreased as the solvent was enriched with EtOH. The dissociation and transfer thermodynamic parameters were calculated, and the results showed that a non‐spontaneous free‐energy change (ΔdissGo>0) and unfavorable enthalpy (ΔdissHo>0) and entropy (ΔdissSo<0) changes occurred on dissociation of trazodone hydrochloride. The free‐energy change or pKa varied nonlinearly with the reciprocal dielectric constant, indicating the inadequacy of the electrostatic approach. The dissociation equilibria are discussed on the basis of the standard thermodynamics of transfer, solvent basicity, and solute‐solvent interactions. The values of ΔtransGo and ΔtransHo increased negatively with increasing EtOH content, revealing a favorable transfer of trazodone hydrochloride from H2O to EtOH/H2O mixtures and preferential solvation of H+ and trazodone (TZD). Also, ΔtransSo values were negative and reached a minimum, in the H2O‐rich zone that has frequently been related to the initial promotion and subsequent collapse of the lattice structure of water. The pKa or ΔdissGo values correlated well with the Dimroth‐Reichardt polarity parameter ET(30), indicating that the physicochemical properties of the solute in binary H2O/organic solvent mixtures are better correlated with a microscopic parameter than the macroscopic one. Also, it is suggested that preferential solvation plays a significant role in influencing the solvent dependence of dissociation of trazodone hydrochloride. The solute‐solvent interactions were clarified on the basis of the linear free‐energy relationships of Kamlet and Taft. The best multiparametric fit to the Kamlet‐Taft equation was evaluated for each thermodynamic parameter. Therefore, these parameters in any EtOH/H2O mixture up to 80% were accurately derived by means of the obtained equations.  相似文献   

17.
A labeling-free surface plasmon resonance (SPR) sensor technique was used to monitor the conformational changes of immobilized globular proteins (RNase A and Lysozyme) in chemical unfolding and refolding. The conformational changes of proteins at solid/liquid interface are characterized as two-state transformation (S-shaped) curves through matrix-effect correction and theoretic estimation. By extrapolation with a Santoro-Bolen equation, the SPR results for both reductive immobilized proteins are estimated to 1.9 kcal mole−1 global free energy (ΔGU) in urea-induced unfolding. But the ΔGU for RNase A and Lysozyme in GdmCl-induced unfolding are 1.5 and 2.15 kcal mole−1, respectively. The disagreement in free energy is partially accounted for by the differences of intra-molecular interactions and immobilization.  相似文献   

18.
The diad tacticity of poly(isopropyl acrylate) was measured from the β-proton absorptions of poly(isopropyl acrylate-α,β-d2) obtained with a 100 MHz NMR spectrometer, and temperature dependence of the tacticity of the polymers obtained by radical polymerization was determined. Enthalpy and entropy differences between isotactic and syndiotactic addition for poly(isopropyl acrylate) were calculated to give the following values: Δ(ΔS) = 0.7 eu; Δ(ΔH) = 0.51 kcal/mole. In the hydrolysis of poly(isopropyl acrylate-α,β-d2), it was found that the rate of hydrolysis of poly(isopropyl acrylate) was dependent on the molecular weight rather than on the tacticity. As for the rate of racemization during hydrolysis, the rate for syndiotactic polymer was much faster than that for the isotactic polymer. The exchange reaction of deuterium at α-position with hydrogen occurred in all the polymers during hydrolysis reaction.  相似文献   

19.
Thermodynamic properties and equilibrium constant of reaction in nanosystems were analyzed theoretically. The effects of sizes of nano-CuO on thermodynamic properties and equilibrium constant were studied using the reaction of nano-copper oxide and sodium bisulfate as a system. The experimental results indicate that with the sizes of reactant decreasing, the molar Gibbs free energy (ΔrGm), the molar enthalpy (ΔrHm) and the molar entropy (ΔrSm) decrease, but the equilibrium constant (K) increases and there are linear trends between the reciprocal of sizes for nano-CuO and the values of ΔrGm, ΔrHm, ΔrSm and Ln K, which are in agreement with the theoretical analysis.  相似文献   

20.
The surface activity and thermodynamic properties for eight low molecular weight nonionic co‐polyester (PE) surfactants have been investigated. Surface and interfacial tensions (IFT) of surfactants in aqueous solutions were measured using the spinning drop technique. From these measurements, the critical micelle concentration (CMC), the surface pressure at CMC (YCMC), the maximum surface concentration (Γmax), the minimum area/molecule at the aqueous solution/air interface (Amin), the effectiveness of surface tension reduction (ΠCMC), the alkane carbon number (nmin) and the IFT at nmin (Ymin) were determined. The thermodynamic parameters of micellization (ΔGmic, ΔHmic and ΔSmic) and of adsorption (ΔGad, ΔHad and ΔSad) for these polymeric surfactants were also calculated. Structural effects on micellization and adsorption are discussed in terms of these parameters. The results show that the ΔGad values were more negative than ΔGmic values for these compounds, so that they favored adsorption before the micellization process. They exhibited IFT in the order of 10−3 to 10−4 mN/m against the thin alkane carbon number range 6–9. This range seemed to be prefered for enhanced oil recovery. Copyright © 2000 John Wiley & Sons, Ltd.  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