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 62 毫秒
1.
The interpretation of studies of dielectric relaxation in polymers is often restricted by ignorance of the mode of motion responsible for dipole orientation in an applied field. Relevant information can be drawn from studies of visco-elastic relaxation, since the visco-elastic relaxation time is comparable with the dielectric relaxation time when the latter requires some ‘whole molecule’ mode of motion involving simultaneous movement of many chain segments. The dielectric and visco-elastic relaxation times are expected to differ when the former involves a localised segmental motion. The dielectric relaxation process is shown to involve a localised segmental motion in acrylic polymers, but involves a ‘whole molecule’ rotation in low molecular weight poly(N-vinylcarbazole). Poly(p-substituted phenyl acetylenes) exhibit two dielectric relaxation processes. A low frequency relaxation comparable with the visco-elastic relaxation is ascribed to the backbone double bonds being polarised unidirectionally along the chain, and a high frequency process is ascribed to a localised segmental motion of the substituted phenyl moiety.  相似文献   

2.
Yuji Hirose 《Polymer》2005,46(6):1913-1920
We report the dielectric properties of poly(styrene oxide)s (PSO) in bulk and concentrated solution states. Since the structure of PSO is asymmetric along the backbone, the repeat unit of PSO is expected to possess the non-zero component of the dipole moment pA parallel to the chain contour as well as the perpendicular component pB. The former and latter cause the dielectric normal mode and segmental mode relaxations, respectively. Contrary to the above mentioned expectation the temperature dependence of ε″ exhibits only the primary (α) relaxation and weak secondary relaxation (β) in the glassy state. No loss peak due to the normal mode relaxation was observed in the frequency region expected from the viscoelastic terminal relaxation in bulk and toluene solutions. The dielectric behaviours of the α relaxation in the bulk state were analyzed in detail and the parameters of the Vogel-Fulcher and the Havriliak-Negami equations were determined. The Kirkwood correlation factor was determined to be 0.36. The 13C NMR spectra indicate that the present PSO samples contain about 2% head-to-head linkages. This cannot be the origin of the disappearance of the normal mode. We conclude that pA of PSO is too small to be detected. The pA calculated with molecular orbital methods supports this conclusion.  相似文献   

3.
DSC, dielectric relaxation and dynamic mechanical thermal analysis (DMTA) were carried out on two multi‐component poly(ether‐co‐amide) samples having different weight ratios of polyamide prepared by condensation polymerization with 12‐aminododecanoic acid, adipic acid and polyetherdiamine consisting of poly(tetramethylene oxide) and poly(propylene oxide). The melting temperature was lowered by an increase in the weight ratio of the polyamide segment. Three relaxation modes, α′, αs and β, were found from dielectric relaxation measurements in different temperature ranges. The high temperature relaxation mode, α′, has a large dielectric constant, which disappears at the melting temperature of the polyamide crystal in the sample. The relaxation times for the segmental motion, αs, were different for the samples, which is attributed to the difference in the composition of the uncrystallized polyamide segments in the amorphous domain. The glass transition temperature estimated from DMTA is located between those of constituting polymers. On the other hand, the activation energy of β‐mode observed at low temperatures is the same for samples with different polyamide ratios, which is attributed to the local motion of the polyether segments. The uncrystallized polyamide segments are miscible with the polyether segments, which results in a lowering of the glass transition temperature of the amorphous domain and enlarges the temperature range of the rubber state of the copolymer due to the high melting temperature of the polyamide segments. © 2016 Society of Chemical Industry  相似文献   

4.
Yu Bian 《Polymer》2009,50(6):1541-1315
Segmental and normal mode dynamics in polyhedral oligomeric silsesquioxane (POSS)/poly(propylene oxide) (PPO) non-reactive and reactive nanocomposites were investigated using a broadband dielectric relaxation spectroscopy (DRS) over wide ranges of frequency and temperature. Three POSS reagents with varying side chain architecture were selected for the study: OctaGlycidyldimethylsilyl (OG), TrisGlycidylEthyl (TG) and MonoGlycidylEthyl (MG). Spectra of OG and TG show a segmental (α) process at lower frequency and a local (β) relaxation at higher frequency, while MG displays only a local relaxation. Neat PPO has both segmental and normal mode (αN) process. In POSS/PPO non-reactive nanocomposites, the presence of OG and TG causes a decrease in the time scale of αN and α relaxation, while MG has no impact on the dynamics of PPO. Chemical reactions in POSS/PPO reactive nanocomposites lead to the formation of nanonetworks. Prior to the onset of reaction, POSS nanoparticles promote the motions of PPO chains, decrease the time scale of relaxation and give rise to thermodielectrically simple spectra. During the reaction, however, the network formation leads to spectral broadening and a gradual increase in the time scale of both segmental (α) and normal mode (αN) relaxation. A detailed account of the effects of structure, concentration and dispersion of POSS in the matrix, molecular weight of PPO, extent of reaction and temperature on the molecular origin, temperature dependence and spectral characteristics of relaxation processes in POSS/PPO nanocomposites is provided.  相似文献   

5.
Fractionated samples of d,l-poly(lactic acid) (PLA) were prepared and the dielectric normal mode relaxation was studied for dilute and semi-dilute solutions of the PLA in a good solvent benzene. Results indicate that in the dilute regime the normal mode relaxation time is proportional to [η]Mw in agreement with the Rouse-Zimm theory, where [η] and Mw denote the intrinsic viscosity and weight average molecular weight, respectively. The dielectric relaxation strength which is proportional to the mean square end-to-end distance 〈r2〉 increases with increasing Mw with the power of 2ν, where ν is the excluded volume parameter determined from [η]. The relaxation time in the semi-dilute regime increases with increasing concentration C due to increases of the entanglement density and the friction coefficient. The relaxation time corrected to the iso-friction state agrees approximately with the dynamic scaling theories. The relaxation strength decreases with increasing concentration indicating that 〈r2〉 decreases on account of the screening of the excluded volume effect. The concentration dependence of 〈r2〉 agrees approximately with the scaling theory proposed by Daoud and Jannink.  相似文献   

6.
Dielectric constant and dielectric loss have been studied for poly(tetrafluoroethylene–perfluorovinyl ether) (PFA) films over a wide temperature range in the frequency range 0.1–100 kHz. Two relaxation peaks were observed, one at room temperature (αa-relaxation) and the other in the range 170–140 K (β-relaxation), with activation energies of 143·2 and 16·4 kcal/mol, respectively. The β-absorption is attributed to the short segmental local mode motion of the main chains. The αa-relaxation can be interpreted as due to large-scale conformational rearrangement. The Cole–Cole diagrams are given at different temperatures and the distribution parameters (ϵ0–ϵ) and (1–α) of the relaxation times were calculated. The X-ray diffraction pattern of PFA shows both a diffuse halo and sharp reflections, characteristic of amorphous and crystalline phases of conventional semicrystalline polymers. Also, no evidence of crystallinity in the films due to thermal treatment during dielectric measurements was observed. IR spectra revealed the absence of any new peaks after the heat treatment.  相似文献   

7.
The proton spin-lattice relaxation and dielectric relaxation were studied in some polymers at the solid–polymer interface was constructed from several filled polymers. A useful model of surface layer which can be considered as consisting of a great number of small solid particles covered with a polymer layer. The following systems were studied: polystyrene, poly(methyl methacrylate), their copolymers and cellulose acetate in the presence of different content of fine particles of aerosil and Teflon. It was established that the decrease of surface layer thickness shifts the minimum of spin-lattice relaxation time T1 of high temperature process to higher temperature and minimum T1 of low temperature process to lower temperature. The same was found for dielectric losses reflecting the motion of side groups and of segments. From temperature dependence of T1 and tan δ for both relaxation processes the apparent energies of activation were calculated. On the base of dielectric relaxation data the circular diagram of complex dielectric constant was constructed and by the Cole-Cole method the dispersion parameter α for polymers at the interface was calculated. These data also show the broadening of relaxation spectra in surface layers. The results are discussed in terms of the restriction of possible conformation of chains at the interface and their interaction with surface. It was established that character of molecular motion changes at the interface is dependent on the mode of molecular motion.  相似文献   

8.
Sumod Kalakkunnath 《Polymer》2006,47(20):7085-7094
The dynamic mechanical and dielectric relaxation properties of a commercial poly(trimethylene terephthalate) [PTT] have been investigated for both quenched and isothermally melt-crystallized specimen films. The relaxation characteristics of PTT were consistent with those of other low-crystallinity semiflexible polymers, e.g. PET and PEEK. While the sub-glass relaxation was largely unperturbed by the presence of the crystalline phase, both calorimetric and broadband dielectric measurements across the glass transition indicated the existence of a sizeable rigid amorphous phase (RAP) fraction in melt-crystallized PTT owing to the constraining influence of the crystal surfaces over the crystal-amorphous interphase region. A strong increase in measured dielectric relaxation intensity (Δ?) with temperature above Tg indicated the progressive mobilization of the RAP material, as well as an overall loss of correlation amongst the responding dipoles.  相似文献   

9.
The effects of concentration and temperature on dynamics of poly(N-isopropylacrylamide)-graft-poly(ethylene oxide) (PNIPAM-g-PEO) chains in semi-dilute aqueous solution were studied by static and dynamic laser light scattering. The intensity-intensity time correlation function shows a fast and a slow relaxation mode, with line widths Γf and Γs, respectively. Γf is scaled to the scattering vector (q) as Γf ∝ q2, revealing that it is due to the cooperative diffusion of the subchains between two neighboring entangled points. As the concentration increases, the slow relaxation becomes slower and contributes more to the total scattered light intensity, indicating that the slow relaxation is related to the chain entanglement. On the other hand, when the solution temperature increases, the PNIPAM chain backbone shrinks, but the fast relaxation remains and the slow mode slows down with a minimum rate at ∼33 °C. It indicates that the slow mode arises from inter-chain clustering, which is gradually suppressed by the intra-chain shrinking. The sample position independence of the time-averaged scattered light intensity 〈IT reveals that the solution is homogeneous and the clustering is transient.  相似文献   

10.
A. Flores  G. Di Marco  M. Pieruccini 《Polymer》2011,52(14):3155-3162
The changes induced in the amorphous regions of poly(ethylene terephthalate) as a consequence of recrystallization processes, taking place after cold-crystallization at Tc = 100 °C, are analyzed by means of isothermal dynamical mechanical spectroscopy and microindentation hardness. Overall, a recrystallization process at either 115 or 125 °C causes an increase of the rigidity within the amorphous domains confined by the crystals. Microhardness measurements carried out at room temperature reveal that recrystallization leads to an enhanced mechanical performance of the amorphous regions. The analysis of isothermal segmental relaxation patterns recorded in a frequency interval of 10−3-60 Hz indicates the appearance of two distinct contributions, which find correspondence with observations by broad band dielectric spectroscopy on the same systems. The faster one is ascribed to segmental relaxation within the amorphous domains where the confinement by crystals is relatively weak. The slow relaxation mode is associated with regions where the conformational dynamics is strongly restricted by the crystals. A relative increase of the slow process is detected upon recrystallization. A recently developed relaxation function model is employed to estimate the size of the static cooperatively rearranging regions for both, the slow and the fast modes. It is found that this size increases either upon decreasing the temperature or as an effect of recrystallization. In addition, the number of monomers involved in a conformational rearrangement turns out to be significantly larger in the regions associated to the slow mode process.  相似文献   

11.
S.J. Rigby  D. Dew-Hughes 《Polymer》1974,15(10):639-644
Dielectric measurements of samples of polymers containing phenylene rings in the back-bone chain have been obtained in the frequency range 62Hz to 100kHz and at temperatures between 4·2K and 330K. Most of the polymers with a phenylene ring in the main chain exhibit a dielectric relaxation in the temperature range 120 to 185K (62Hz). In poly(2,6-dimethyl-p-phenylene oxide) a single relaxation region occurs, ranging from 125K (62Hz) to 170K (100kHz). In poly(phenylene sulphide) a very broad relaxation region occurs over the temperature range ~70K to 190K. Poly(p-xylylene) exhibits a well defined relaxation centred at 185K (62Hz) moving to 250K (100kHz). In poly(dichloro-p-xylylene) the relaxation region is in the same temperature range as for poly(p-xylylene), although less sharp; in poly(monochloro-p-xylylene) the beginnings of relaxation region are visible at 330K (62Hz). In all these polymers the low temperature relaxation region is believed to be caused by a local re-orientation of the phenylene ring, or substituted phenylene ring. The substitution of a single chlorine atom on the phenylene ring causes the relaxation to appear at higher temperatures, whereas the disubstitution causes the relaxation region to re-appear at a similar temperature to that for the unsubstituted polymer. The activation energies of these processes lie in the range 30–45MJ/kg mol (7–11 kcal/mol). The temperatures and activation energies of this process appear to depend on the units adjacent to the phenylene ring, and on steric and polar effects caused by substituents on the ring.  相似文献   

12.
Sanja Risti? 《Polymer》2008,49(21):4695-4702
An investigation was carried out on the molecular dynamics of blends composed of poly(amidoamine) (PAMAM) dendrimers with ethylenediamine core and amino surface groups (generations 0 and 3) and three linear polymers: poly(propylene oxide) - PPO and two block copolymers, poly(propylene oxide)/poly(ethylene oxide) - PPO/PEO with different mole ratios: 29/6 (amorphous) and 10/31 (crystalline). The results were generated over a broad range of frequency and temperature by Dielectric Relaxation Spectroscopy (DRS) and Dynamic Mechanical Spectroscopy (DMS). Dielectric spectra of dendrimers in the PPO matrix reveal a decrease in the time scale of normal and segmental relaxation with increasing dendrimer concentration. In the amorphous blends with 29PPO/6PEO matrix, no effect of concentration on the time scale of normal and segmental processes was observed. But in the crystalline blends with 10PPO/31PEO matrix, relaxation time increases with increasing dendrimer concentration. Results acquired by DRS and DMS were contrasted and the obtained relaxation times were found to be in excellent agreement. A detailed analysis of the effect of generation and concentration of dendrimers, hydrophilicity and morphology of the polymer matrix and temperature on the molecular origin, the shape of the relaxation spectra, the dielectric relaxation strength and the frequency location for the maximum loss in dendrimer-polymer blends is provided.  相似文献   

13.
A novel rigid fluorescent probe, carbazole-terephthalate cyclophane (Cz-TP) was applied to evaluate local dielectric constants (ε) of various polymer solids in a wide range of temperatures. For poly(vinylidene fluoride), the ε increased above the glass transition temperature (Tg), due to relaxations of the polar segment -(CH2CF2)- of the main chain. For poly(alkyl methacrylate)s, the ε increased above the Tg or the melting temperature of the side chain, where motions of the polar ester groups are activated. For cyanoethylated polymers, the ε increased owing to motions of the polar cyano groups at the end of the side chain and the ε corresponded to the dielectric constant evaluated by dielectric relaxation measurement at a high frequency, because the Cz-TP exciplex has a lifetime of tens of nanoseconds. For a cyanoethylated polymer with a high content of cyano groups, the ε was larger at low temperatures than the dielectric constant obtained by the macroscopic dielectric relaxation measurement. These results show that the Cz-TP molecule is a useful probe for evaluation of the local polarity in polymer solids over a wide temperature range and can detect even a small change in ε at transition temperatures such as glass transition, side-chain melting, and side-chain relaxation.  相似文献   

14.
The polymer nanocomposite (PNC) films consisted of poly(ethylene oxide) (PEO) and poly(methyl methacrylate) (PMMA) blend matrices dispersed with nanoparticles of amorphous silica (SiO2) have been prepared by solution-cast method followed by melt-press technique. Effects of SiO2 concentration (x?=?0, 1, 3 and 5 wt%) and PEO–PMMA blend compositional ratios (PEO:PMMA?=?75:25, 50:50, and 25:75 wt%) on the surface morphology, crystalline phase, polymer-polymer and polymer-nanoparticle interactions, melting phase transition temperature, dielectric permittivity, electrical conductivity, electric modulus and the impedance properties of the PNC films have been investigated. The crystalline phase of the PNC films decreases with the increase of PMMA contents which also vary anomalously with the increase of SiO2 concentration in the films. The melting phase transition temperature and polymer-nanoparticle interactions significantly change with the variation in the compositional ratio of the blend polymers in the PNC films. It is observed that the effect of SiO2 on the dielectric and electrical properties of these PNCs vary greatly with change in the compositional ratio of PEO and PMMA in the blends. The dielectric relaxation process of these films confirm that the polymers cooperative chain segmental dynamics becomes significantly slow when merely 1 wt% SiO2 nanoparticles are dispersed in the polymer blend matrix.  相似文献   

15.
Blends of poly(ether ether ketone) (PEEK) and poly(aryl ether sulphone) (PES) have been prepared in the whole composition range. The molecular dynamics and α-relaxation behaviour of these materials have been studied using dynamic mechanical and dielectric relaxation spectroscopy. From dynamic mechanical relaxation studies, two α-relaxation peaks corresponding to the segmental relaxation process of pure components in the blend was observed. Also, it was found that the temperature at which α-process of the homopolymers occurs, shows a slight change with blend composition, corresponding to a PEEK-rich and PES-rich phase. The relaxation intensities of the homopolymers in the blend compared to that in pure state were approximately proportional to their respective content in the blend. From the phase composition of the respective phases obtained using Fox equation, it has been inferred that PEEK dissolves more in PES than vice-versa. The α-relaxation of PES could not be detected from dielectric relaxation spectroscopy because of the possible influence of dc conduction and electrode polarization losses. Otherwise, the α-relaxation behaviour of PEEK-rich phase observed from dielectric relaxation studies agree with those inferred from dynamic mechanical relaxation studies. Furthermore, activation energies for molecular motions (Ea) at the α-relaxation have also been determined using an Arrhenius form of equation and it has been found that Ea for both PEEK-rich and PES-rich phase show variation with composition. Similarly, the relaxation times associated with the mobility of relaxing species in both PEEK and PES are influenced in the blends. It is likely that these observations are related to some interactions and a partial segmental mixing between the blend components, which result in changes in the local molecular environment on blending.  相似文献   

16.
A series of dangling chain based-polyurethane/poly(methyl methacrylate) (DPU/PMMA) filled with exfoliated layered double hydroxides (LDH) were synthesized by methyl methacrylate in-situ intercalative polymerization. The dangling chains were introduced by using vegetable oils as chain extender. The effect of dangling chain and the contents of LDH on the molecular dynamics of DPU/PMMA was investigated by a combination of dynamic mechanical analysis and broadband dielectric relaxation spectroscopy. Compared with polyurethane/poly(methyl methacrylate) (PU/PMMA) without dangling chain, the glass transition temperature (Tg) of DPU/PMMA shifts to lower temperature and the segmental dynamics becomes faster. A plateau with a high loss factor value above Tg significantly broadens the damping temperature range due to the synergy effect between the dangling chains and LDH layers. In DPU/PMMA/LDH nanocomposites, the α-relaxation associated with the glass transition of the polymer matrix becomes slower with the increase of LDH content, which indicates a restricted molecular mobility in the interfacial regions between polymer and LDH. However, the local relaxations at relatively low temperature remain unaffected by dangling chain or the addition of LDH. When the LDH content increases, Maxwell–Wagner–Sillars (MWS) interfacial polarization process caused by charge accumulation at interfaces becomes faster because of the smaller mean distance d between the exfoliated LDH layers.  相似文献   

17.
Proton spin-lattice relaxation times T1 were measured for three poly(dimethylsiloxanes) of different molecular weight over the temperature range 3K–300K. The T1 temperature dependence is characterized by different relaxation mechanisms due to hindered methyl group rotation, segmental motion about the SiO bond, a motional process associated with recrystallization and molecular motions which are effective in the temperature range below the methyl group rotational minimum. None of these processes showed a molecular weight effect within the range covered by the three samples. A temperature dependent paramagnetic impurity effect was observed. It can be concluded from the experimental results that this effect is due to molecular oxygen.  相似文献   

18.
Broad band dielectric relaxation spectra are reported on a range of polymers created by varying the ether segment in a series of poly(ether imide)s. Changes in the structure allow the effects of steric constraints on the local conformational dynamics of the polymer chain to be explored. These changes have a significant effect on the glass transition temperatures of these polymers which range from 245 to over 420°C. In contrast, the low temperature dielectric relaxation behavior of these polymers is very similar and is attributed to cooperative local oscillatory—librational motions. Changes in the stereochemistry effect the amplitude, activation energy for the relaxation process, the packing chain density, and values of the high frequency limiting permittivity, ε′. This latter parameter is sensitive to the extent of dipole induced dipole and π‐π electron interactions and is influenced by the packing density. The magnitude of ε′ is a very important parameter in determining the suitability of poly(imide)s for electrical applications. The magnitude of ε′ increases with the density; however, deviations from this general trend are observed when large nonpolar groups inhibit the interaction of neighboring chains. © 2014 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2014 , 131, 41191.  相似文献   

19.
The molecular relaxation characteristics of rubbery amorphous crosslinked networks based on poly(ethylene glycol) diacrylate [PEGDA] and poly(propylene glycol) diacrylate [PPGDA] have been investigated using broadband dielectric spectroscopy. Dielectric spectra measured across the sub-glass transition region indicate the emergence of an intermediate “fast” relaxation in the highly crosslinked networks that appears to correspond to a subset of segmental motions that are more local and less cooperative as compared to those associated with the glass transition. This process, which is similar to a distinct sub-Tg relaxation detected in poly(ethylene oxide) [PEO], may be a general feature in systems with a sufficient level of chemical or physical constraint, as it is observed in the crosslinked networks, crystalline PEO, and PEO-based nanocomposites.  相似文献   

20.
Two series of hyperbranched polymers (HP), polyurethanes and polyureas, with aromatic and aliphatic structures, are synthesized in one-pot method using commercially available monomers. The obtained HP samples were characterized by 1H Nuclear Magnetic Resonance (NMR) spectroscopy, Gel Permeation Chromatography (GPC), Differential Scanning Calorimetry (DSC) and Thermal Gravimetric Analysis (TGA) measurements. Molecular dynamics in these systems were investigated by combining Thermally Stimulated Depolarization Currents (TSDC) and broadband Dielectric Relaxation Spectroscopy (DRS) techniques. High conductivity contribution in dielectric loss does not allow the study of the segmental α relaxation associated with the glass transition. In the glassy state two secondary relaxation mechanisms have been investigated, the γ and the β mechanism. The γ relaxation mechanism, at low temperatures/high frequencies, is attributed to motions of the end groups (-OH for polyurethanes and -NH2 for polyureas), and has been found faster in the hyperbranched polyureas. In addition, our results reveal that γ relaxation mechanism in both series depends on the chemical structure, being faster for aliphatic structures. The β relaxation mechanism, at higher temperatures/lower frequencies, is attributed to the motions of branched ends with polar groups. Our study suggests that this mechanism may be a typical relaxation process for hyperbranched polyurethanes structures, not existed in the linear counterparts. All the systems exhibit dc conductivity at temperatures higher than Tg which shows Arrhenius-like temperature dependence and is characterized by rather high activation energies (in the order of 200 kJ/mol). At temperatures lower than Tg all the systems studied exhibit remarkably high charge mobility. In particular, aliphatic hyperbranched polyureas exhibit dc conductivity which has been found to be of VTF type concerning the temperature dependence. This result implies that the conduction mechanism is coupled with molecular motions in the glassy state of the polymer.  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