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 93 毫秒
1.
Inverse gas chromatography (IGC) has been widely used to determine the Flory–Huggins parameters, χ, between solutes (probes) and polymers. This study correlated the Flory–Huggins parameter data using a multiparameter model, which included dispersion, polarity, acidity, and basicity components. The parameters of poly(ε‐caprolactone) (PCL) and polyepichlorohydrin (PECH) were calculated from IGC data using a series of probes. The parameters of the polymers were used to evaluate mutual miscibility between PCL and PECH. The results predicted miscibility in agreement with the conclusion of an IGC study using blends of PCL and PECH. A method to estimate the confidence interval of polymer parameters was proposed. The anomalous solubility parameter of polymer mixtures previously reported was also explained using this model. © 2009 Wiley Periodicals, Inc. J Appl Polym Sci, 2009  相似文献   

2.
Amylopectin (AP), a potato‐starch‐based polymer with a molecular weight of 6,000,000 g/mol, was blended with poly(ε‐caprolactone) (PCL) and characterized with inverse gas chromatography (IGC), differential scanning calorimetry (DSC), and X‐ray diffraction (XRD). Five different compositions of AP–PCL blends ranging from 0 to 100% AP were studied over a wide range of temperatures (80–260°C). Nineteen solutes (solvents) were injected onto five chromatographic columns containing the AP–PCL blends. These solutes probed the dispersive, dipole–dipole, and hydrogen‐bonding interactions, acid–base characteristics, wettability, and water uptake of the AP–PCL blends. Retention diagrams of these solutes in a temperature range of 80–260°C revealed two zones: crystalline and amorphous. The glass‐transition temperature (Tg) and melting temperature (Tm) of the blends were measured with these zones. The two zones were used to calculate the degree of crystallinity of pure AP and its blends below Tm, which ranged from 85% at 104°C to 0% at Tm. IGC complemented the DSC method for obtaining the Tg and Tm values of the pure AP and AP–PCL blends. These values were unexpectedly elevated for the blends over that of pure AP and ranged from 105 to 152°C for Tg and from 166 to 210°C for Tm. The Tm values agreed well with the XRD analysis data. This elevation in the Tg and Tm values may have been due to the change in the heat capacity at Tg and the dependence of Tg on various variables, including the molecular weight and the blend composition. Polymer blend/solvent interaction parameters were measured with a variety of solutes over a wide range of temperatures and determined the solubility of the blends in the solutes. We were also able to determine the blend compatibility over a wide range of temperatures and weight fractions. The polymer–polymer interaction coefficient and interaction energy parameter agreed well on the partial miscibility of the two polymers. The dispersive component of the surface energy of the AP–PCL blends was measured with alkanes and ranged from 16.09 mJ/m2 for pure AP to 38.26 mJ/m2 when AP was mixed with PCL in a 50/50% ratio. This revealed an increase in the surface energy of AP when PCL was added. © 2006 Wiley Periodicals, Inc. J Appl Polym Sci 101: 3076–3089, 2006  相似文献   

3.
Experimental values of the Flory–Huggins parameter, χ, between polymers and solvents, are frequently used to determine the solubility parameters of the polymers. A method using nonlinear curve fitting of RTχ/V was compared to the linear regression method commonly used. It was found that the formulas for the solubility parameter were the same, but the linear method produced a slightly different entropy term. The nonlinear method gave a lower correlation coefficient and wider confidence intervals and was more effective at distinguishing systems than the linear model. The effect of the deviation of probes in the solubility parameter model is discussed. Using probes with low solubility parameters to measure the polymer solubility parameter gave wider confidence intervals. © 2004 Wiley Periodicals, Inc. J Appl Polym Sci 91: 2894–2902, 2004  相似文献   

4.
A formulation for the solubility parameter based on the group‐contribution, lattice fluid equation of state was derived. The solubility parameters of pure liquid solvents, polymers, copolymers, and liquid mixtures were calculated and compared against the best available data. This investigation was conducted on pure components and mixtures of alkanes, alkenes, ketones, ethers, acetates, alcohols, chlorinated molecules, and cyclic and aromatic solvents. The capabilities of the model to distinguish between two isomers and to predict the solubility parameter of supercritical fluids and their mixtures were also studied. The predicted values are generally good, although the error increases when hydrogen bonding is present. A primary application of the procedure is for the prediction of the solubility parameters of polymers. © 2001 John Wiley & Sons, Inc. J Appl Polym Sci 80: 197–206, 2001  相似文献   

5.
In a recent study, a two‐dimensional solubility parameter model was used to correlate the heat of solution for solutes ranging from n‐alkanes to alcohols, dissolved in isotatic polypropylene (PP), poly(ethyl ethylene) (PEE), and poly(dimethylsiloxane) (PDMS). When literature data of solubility parameter components of solutes were used, the correlation had some scattering for solutes with low values of cohesive energy density. In this study, the components of solubility parameters of solutes and polymers were estimated from cohesive energy and heat of sorption of solutes. Good correlation was obtained for the specific heat of sorption (ΔUsorp/V) for solutes ranging from n‐alkanes to alcohols, and PDMS had a polar component as previously estimated. Free volume effect in solution process may be the source of a small systematic deviation from the model. © 2009 Wiley Periodicals, Inc. J Appl Polym Sci, 2009  相似文献   

6.
Films of poly(ε‐caprolactone) (PCL) blends with α‐chitin and chitosan were prepared as completely biodegradable polyester/polysaccharide composites. DSC thermal analysis revealed that the crystallization of PCL was suppressed by blending with α‐chitin and chitosan. The specific interaction between PCL and polysaccharides was investigated by FTIR spectroscopy. The PCL carbonyl vibration bands could be resolved into three components: crystalline, amorphous and interacting. The FTIR spectra indicated that there were hydrogen bonding interactions between PCL and polysaccharides, and that polysaccharides suppressed the crystallization of PCL, consistent with the results obtained by DSC. © 2001 Society of Chemical Industry  相似文献   

7.
Different hydroxyl content poly(styrene‐cop‐(hexafluoro‐2‐hydroxylisopropyl)‐α‐methylstyene) [PS(OH)‐X] copolymers were synthesized and blends with 2,2,6,6‐tetramrthyl‐piperdine‐1‐oxyl end spin‐labeled PEO [SLPEO] were prepared. The miscibility behavior of all the blends was predicted by comparing the critical miscible polymer–polymer interaction parameter (χcrit) with the polymer–polymer interaction parameter (χ). The micro heterogeneity, chain motion, and hydrogen bonding interaction of the blends were investigated by the ESR spin label method. Two spectral components with different rates of motion were observed in the ESR composite spectra of all the blends, indicating the existence of microheterogeneity at the molecular level. According to the variations of ESR spectral parameters Ta, Td, ΔT, T50G and τc, with the increasing hydroxyl content in blends, it was shown that the extent of miscibility was progressively enhanced due to the controllable hydrogen bonding interaction between the hydroxyl in PS(OH) and the ether oxygen in PEO. © 2004 Wiley Periodicals, Inc. J Appl Polym Sci 92: 2312–2317, 2004  相似文献   

8.
Diblock copolymers with different poly(ε‐caprolactone) (PCL) block lengths were synthesized by ring‐opening polymerization of ε‐caprolactone in the presence of monomethoxy poly(ethylene glycol) (mPEG‐OH, MW 2000) as initiator. The self‐aggregation behaviors and microscopic characteristics of the diblock copolymer self‐aggregates, prepared by the diafiltration method, were investigated by using 1H NMR, dynamic light scattering (DLS), and fluorescence spectroscopy. The PEG–PCL block copolymers formed the self‐aggregate in an aqueous environment by intra‐ and/or intermolecular association between hydrophobic PCL chains. The critical aggregation concentrations of the block copolymer self‐aggregate became lower with increasing hydrophobic PCL block length. On the other hand, reverse trends of mean hydrodynamic diameters were measured by DLS owing to the increasing bulkiness of the hydrophobic chains and hydrophobic interaction between the PCL microdomains. The partition equilibrium constants (Kv) of pyrene, measured by fluorescence spectroscopy, revealed that the inner core hydrophobicity of the nanoparticles increased with increasing PCL chain length. The aggregation number of PCL chain per one hydrophobic microdomain, investigated by the fluorescence quenching method using cetylpyridinium chloride as a quencher, revealed that 4–20 block copolymer chains were needed to form a hydrophobic microdomain, depending on PCL block length. © 2006 Wiley Periodicals, Inc. J Appl Polym Sci 99: 3520–3527, 2006  相似文献   

9.
We have investigated the enhancement in miscibility, upon addition of bisphenol A (BPA) of immiscible binary biodegradable blends of poly(ε‐caprolactone) (PCL) and poly(L ‐lactide) (PLLA). That BPA is miscible with both PCL and PLLA was proven by the single value of Tg observed by differential scanning calorimetry (DSC) analyses over the entire range of compositions. At various compositions and temperatures, Fourier transform infrared spectroscopy confirmed that intermolecular hydrogen bonding existed between the hydroxyl group of BPA and the carbonyl groups of PCL and PLLA. The addition of BPA enhances the miscibility of the immiscible PCL/PLLA binary blend and transforms it into a miscible blend at room temperature when a sufficient quantity of the BPA is present. In addition, optical microscopy (OM) measurements of the phase morphologies of ternary BPA/PCL/PLLA blends at different temperatures indicated an upper critical solution temperature (UCST) phase diagram, since the ΔK effect became smaller at higher temperature (200°C) than at room temperature. An analysis of infrared spectra recorded at different temperatures correlated well with the OM analyses. © 2006 Wiley Periodicals, Inc. J Appl Polym Sci 100: 1146–1161, 2006  相似文献   

10.
Enzymatically prepared novel polyphenol poly(4,4′‐dihydroxydiphenyl ether) (PDHDPE) is blended to modify the properties of biodegradable polyester poly(?‐caprolactone) (PCL). Since the differential scanning calorimetry data show single composition‐dependent glass transition for each blend, PCL and PDHDPE are found to be miscible in the amorphous phase. The crystallization of PCL is depressed by PDHDPE because PDHDPE reduces the molecular mobility and the flexibility of molecular chains of PCL. The Fourier transform infrared spectra clearly indicate that PCL and PDHDPE interact through strong intermolecular hydrogen bonds formed between the carbonyl groups of PCL and the hydroxyl groups of PDHDPE. The increase of the long period, calculated on the basis of Bragg's law with the measurement of small‐angle X‐ray scattering, is found because the peak position of the profiles of Lorentz‐corrected intensity shifts to smaller angle. With the help of lamellar stack model and one‐dimensional correlation function, the accurate lamellar parameters are calculated. The increase of long period is induced by the increase of crystal thickness. The thermal treatment can effectively modify the thermal stability of PCL/PDHDPE blends with the introduction of an intermolecular coupling of the polymer to give crosslinked and/or branched products. It is also found that the addition of PDHDPE to PCL would obviously increase the Young's modulus of PCL. © 2006 Wiley Periodicals, Inc. J Appl Polym Sci 101: 149–160, 2006  相似文献   

11.
A model based on a combination of the solubility parameters of Hansen and the polymer solution theory of Flory–Rehner was used to predict the solubility and permeation properties of organic solvents in Viton polymeric glove material. To test the validity of the model, weight gain data were collected for 32 organic solvents versus Viton. Samples were exposed for periods of 2 to 12 weeks until each glove sample had achieved a stable, equilibrium weight. Using a nonlinear least‐squares regression, the three‐dimensional solubility parameter was determined for Viton to be as follows: dispersion = 15.38, polar = 10.49, and hydrogen bonding = 2.47. Breakthrough times, lag times, and steady‐state permeation rates for the solvents verses Viton were also determined and combined with results obtained from the literature. A high level of correlation was observed between the model and the properties evaluated. © 2004 Wiley Periodicals, Inc. J Appl Polym Sci 93: 2688–2698, 2004  相似文献   

12.
Some hydrogen‐bonding‐interactions‐mediated poly(ε‐caprolactone) (PCL)/silica (SiO2) polymeric‐inorganic hybrids were prepared by the sol–gel process. The non‐isothermal crystallization kinetics of PCL in the hybrids was investigated by means of differential scanning calorimetry. The results show that the Jeziorny method, together with the combination of the Avrami and Ozawa equations, is applicable to describe the non‐isothermal crystallization kinetics of the PCL in the hybrid system, while the Ozawa theoretical method can be used just for the pure PCL and the 70/30 TEOS/PCL hybrids. Depending on the composition of the PCL/silica, the hybrids displayed microphase separation at various scales, which in turn affect the crystallization behavior and mechanism of PCL in the hybrids. On the one hand, the inorganic component (viz SiO2) could act as the nucleating agent to facilitate the crystallization of PCL in the hybrids. On the other hand, the SiO2 networks also confine the crystallization of PCL. Copyright © 2004 Society of Chemical Industry  相似文献   

13.
Optically pure polylactides, poly(L ‐lactide) (PLLA) and poly(D ‐lactide) (PDLA), were blended across the range of compositions with poly(ε‐caprolactone) (PCL) to study their crystallization, morphology, and mechanical behavior. Differential scanning calorimetry and dynamic mechanical analysis (DMA) of the PLA/PCL blends showed two Tgs at positions close to the pure components revealing phase separation. However, a shift in the tan δ peak position by DMA from 64 to 57°C suggests a partial solubility of PCL in the PLA‐rich phase. Scanning electron microscopy reveals phase separation and a transition in the phase morphology from spherical to interconnected domains as the equimolar blend approaches from the outermost compositions. The spherulitic growth of both PLA and PCL in the blends was followed by polarized optical microscopy at 140 and 37°C. From tensile tests at speed of 50 mm/min Young's modulus values between 5.2 and 0.4 GPa, strength values between 56 and 12 MPa, and strain at break values between 1 and 400% were obtained varying the blend composition. The viscoelastic properties (E′ and tan δ) obtained at frequency of 1 Hz by DMA are discussed and are found consistent with composition, phase separation, and crystallization behavior of the blends. POLYM. ENG. SCI., 46:1299–1308, 2006. © 2006 Society of Plastics Engineers  相似文献   

14.
An amphiphilic graft copolymer, hydroxypropylcellulose‐graft‐poly(ε‐caprolactone) (HPC‐g‐PCL), was synthesized by bulk polymerization without a catalyst and characterized with one‐dimensional and two‐dimensional NMR spectroscopy. Molar substitution of ε‐caprolactone on HPC (MSCL) was estimated by both gravimetry and 1H‐NMR, and the gravimetric method was considered suitable for MSCL determination. Heterogeneity in the HPC‐g‐PCL film was suggested by a microscopic study, and the existence of PCL‐rich crystalline regions was confirmed by the results of X‐ray diffraction and differential scanning calorimetry (DSC). The double endotherm observed in the DSC scans of HPC‐g‐PCL was associated with the different molecular weight fractions in the copolymer. © 2003 Wiley Periodicals, Inc. J Appl Polym Sci 89: 718–727, 2003  相似文献   

15.
Poly(?-caprolactone) (PCL) of ca. 20,000 molecular weight is shown to be partially miscible with three aromatic-amine-cured epoxy resins. This conclusion is based on the depression of the epoxy Tg, the effect on physical and mechanical properties, and the observation that a large proportion (40-55%) of the PCL ester groups are involved in hydrogen bonding. This miscibility behavior is compared to PCL blends with anhydride-cured epoxy resins, which appear to have a two-phase morphology. The different miscibilities are rationalized on the basis of the existence of functional groups (e. g., hydroxyl) in amine-cured epoxies which are capable of hydrogen bonding to the PCL ester groups. Anhydride-cured epoxy resins contain fewer potential hydrogen bonding sites.  相似文献   

16.
用溶解度参数法包括三维溶解度参数基团贡献计算法、反气相色谱法以及差示扫描量热(DSC)法研究了聚碳酸酯(PC)和聚(1,4-环己烷二甲酸-1,4-环己烷二甲醇)酯(PCCD)的相客性.研究了色散、极性和氢键作用力对PC和PCCD相容性的影响.用基团贡献计算法得到PC和PCCD的三维溶解度参数非常接近,并且每一维度(即色...  相似文献   

17.
The intermolecular hydrogen bonding interactions between poly(3‐hydroxybutyrate) and poly(styrene‐co‐vinyl phenol) copolymers with mutual solvent epichlorohydrin were thoroughly investigated by steady‐state fluorescence and viscosity techniques. Fluorescence spectroscopy along with viscosity technique was used to asses the intermolecular hydrogen bonding between poly‐(3‐hydroxybutyrate) and its blends with five copolymer samples of styrene–vinyl phenol, containing different proportions of vinyl phenol but similar average molecular weight and polydispersity index. In the case of very low OH contents (2–4 mol %), as expected, both components of poly(3‐hydroxybutyrate) and poly(styrene‐co‐4‐vinylphenol) chains are well separated and remain so independently of the mixed polymer ratio and overall polymer concentration as well. Conversely, when the OH content reaches 5.8 mol % or more, a significant decrease of the intrinsic fluorescence intensity emitted by the copolymer is detected upon addition of aliquots of poly(3‐hydroxybutyrate). In these cases, an average value for the interassociation equilibrium constant, KA = 8.7, was obtained using a binding model formalism. A good agreement of these results with those obtained from complementary viscosity measurements, through the interaction parameter, Δb, was found. © 2006 Wiley Periodicals, Inc. J Appl Polym Sci 100: 900–910, 2006  相似文献   

18.
Inverse gas chromatography has been widely used to determine the Flory–Huggins parameter, χ, between two solvents. Many studies showed that interaction parameters were probe dependent. In recent studies, it was proposed that the interaction between two solvents may lead to different contact probability between solutes and solvent mixtures and create an apparent solubility parameter different from volume average rule. An equation was previously derived to relate the probe dependency to the deviation of solubility parameter from the volume average rule. By plotting ?2?3RT23/V2) versus the solubility parameter of solutes, a linear trend could be observed with a negative slope for miscible mixtures. When there was an unfavorable interaction between two solvents, an opposite situation would be observed. In this study, mixtures of 19,24‐dioctadecyldotetracontane (C78) and its derivatives were tested. The solubility parameters of mixtures showed negative deviation from the volume average. The plots of ?2?3RT23/V2) versus solubility parameter of solutes had positive slopes. For two derivatives the best estimated values of RT23/V2) were negative in certain temperatures. Enthalpy–entropy compensation plot showed that these two derivatives have higher entropy of mixing. © 2012 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2013  相似文献   

19.
Poly(ε‐caprolactone) (PCL) was melt compounded with “Bucky gels”‐like mixture that prepared by grinding multiwalled carbon nanotubes (MWNTs) and ionic liquids (ILs). Raman spectrum showed the significant interaction between ILs and MWNTs. The dielectric behavior of PCL nanocomposites based on unmodified and IL‐modified MWNTs was studied from 40 Hz to 30 MHz. The addition of ILs significantly enhanced the dielectric property of PCL/IL/MWNT ternary nanocomposites, which was much higher than that of the sum of PCL/IL with PCL/MWNT binary nanocomposites. The dielectric properties of PCL/IL/MWNT nanocomposites were mainly influenced by ILs in low frequency and were dominated by MWNTs in high frequency. SEM results revealed that a more uniform and fine dispersion of MWNTs were achieved throughout the PCL matrix because of ILs. The addition of ILs in nanocomposites changed the crystallinity of PCL. © 2013 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2014 , 131, 40231.  相似文献   

20.
The aqueous solubilities (logS) of a set of 27 5‐amino‐1‐aryl‐1H‐tetrazoles with known biological activity were determined photometrically and correlated with their octan‐1‐ol/water partition coefficients (logP). The otained linear relationship between logS and logP can be significantly improved by a melting point term, an electronic term and several indicator variables describing the deviating solubility behaviour of compounds with hydrogen bonding groups and the dependence on substituent position. The calculated moreparameter equations allow the reliable prediction of aqueous solubilities for 5‐amino‐1‐aryl‐1H‐tetrazoles not yet synthesized or investigated.  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