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 78 毫秒
1.
The dynamic compliances J′(ω) and J″(ω) and the dielectric permittivities ?′(ω) and ?″(ω) are reported over a wide range of frequency ω and temperature for comb-branched and brush-shaped poly(n-butyl acrylate) prepared by atom transfer radical polymerization. The analysis here of the viscosity η for the comb- and brush-shaped polymers indicates the need to account for an increase of the persistence length with increasing density of the side chains. Enhanced values of JS are attributed to a dilution effect arising from the side chains on the values that would otherwise arise from the backbone chain. The dielectric loss ?″(ω) demonstrates a deviation from frequency-temperature superposition at a certain range of frequency, with the deviation increasing with increasing density of the side chains. The deviation occurs for a frequency range for which J′(ω) and J″(ω) are approaching their terminal response, but no corresponding deviation from frequency-temperature superposition is noted for these functions. The dielectric behavior in this region is attributed to a δ-relaxation at frequencies lower than the principal α-relaxation, similar to behavior reported for certain polymers with mesogenic side chains.  相似文献   

2.
Jong Keun Park 《Polymer》2009,50(24):5720-6884
Perfluorosulfonate ionomers (PFSIs) neutralized by tetraalkylammonium ions have been investigated using variable temperature 1H and 13C solid-state NMR (ssNMR) spectroscopy to probe the thermally induced properties of tetraalkylammonium ions at temperatures near the α-relaxations. Tetramethylammonium (TMA+), tetraethylammonium (TEA+), tetrapropylammonium (TPA+) and tetrabutylammonium (TBA+) ions have been incorporated in our study for the systematic control of ionic interactions within PFSIs according to the chain-length of ammonium ions. 13C static ssNMR results show that bulkier TPA+ and TBA+ ions undergo molecular tumbling motions near or above the α-relaxation temperature of the corresponding PFSI ionomers, with jumping rates of κ ≈ 1 kHz. Moreover, the results of 1H spin-lattice (T1) relaxation time measurements suggest that smaller TMA+ ions in the PFSI exhibit a thermally induced ordering effect as the sample temperature approaches the α-relaxation temperature of TMA+-PFSI. This ordering phenomenon is also supported by the results from small-angle X-ray scattering.  相似文献   

3.
We report on electrical conductivity relaxation measurements of solid polymer electrolytes (SPE) based on poly(vinyl alcohol) (PVOH) and LiClO4 in which nanoporous Al2O3 particles with average pore diameter of 58 Å were dispersed. A power law frequency dependence of the real part of the electrical conductivity is observed as a function of temperature and composition. This behaviour is typical of systems in which correlated ionic motions in the SPE bulk material are responsible for ionic conductivity. This variation is well fitted to a Jonscher expression σ′(ω) = σ0[1 + (ω/ω0)p] where σ0 is the dc conductivity, ω0 the characteristic angular frequency relaxation and p is the fractional exponent between 0 and 1. For a prototype membrane with composition 0.9PVOH − 0.1LiClO4 + 7 wt.%Al2O3, it was found that the temperature dependence of σ0 and ω0, may be described by the VTF relationship, ? = ?0 exp[−B/(T − T0)], with approximately the same constant B and reference temperature T0, indicating that ion mobility is coupled to the motions of the polymer chains. Moreover, p decreased with increasing temperature, from 0.68 at T = 319 K, to 0.4 at T = 437 K, indicating weaker correlation effects among mobile ions when the temperature is increased.  相似文献   

4.
Wen-Hong Liu  Hsiu-Li Lin 《Polymer》2007,48(14):4152-4165
Using dynamic light scattering (DLS) and capillary dynamic viscoelasticity (DVE) analyzer, we investigated dilute (0.5 mg/ml) poly(diallyl dimethyl ammonium chloride) (PDADMAC) aqueous solution properties for three different molecular weights of PDADMACs mixed with various concentrations of NaCl. The dependence of PDADMAC molecular chain conformations in aqueous solutions on polymer molecular weight and NaCl concentration were studied. By analyzing dynamic shear viscosity η′(ω), viscoelastic relaxation times tr, and shear rate at tube wall ?a(ω) of PDADMAC aqueous solutions in oscillatory flows, we proposed that polymer chain conformations varied with increasing shear frequency ω via the following steps: intra-polymer associations, dissociation of intra-polymer associations, stretching of polymer chains, inter-polymer aggregations, and dissociations of inter-polymer aggregations. The intra-polymer associations lowered the n′ exponent of storage modulus G′(ω) (G′(ω) ∼ ωn) with n′ < 2, and the polymer chain stretching and inter-polymer aggregations caused shear thickening (i.e. upturn of η′(ω)) of PDADMAC aqueous solutions. The behaviors of the lowering of n′ exponent with n′ < 2 and the shear thickening were favored by increasing ionic strength of solutions. By comparing η′(ω) data with DLS hydrodynamic radii (Rh) data, we also confirmed the possibility of inter-polymer aggregations in dilute solutions when polymer chains were stretched in oscillatory flows.  相似文献   

5.
Kumiko Asai  Kohji Tashiro 《Polymer》2008,49(19):4298-4306
To understand the effect of the nano-filler particles on the crystallization kinetics and crystalline structure of poly(vinylidene fluoride) (PVDF) upon nano-composite formation, we have prepared PVDF/organically modified layered titanate nano-composite via melt intercalation technique. The layer titanate (HTO) is a new nano-filler having highly surface charge density compared with conventional layered silicates. The detailed crystallization behavior and its kinetics including the conformational changes of the PVDF chain segment during crystallization of neat PVDF and HTO-based nano-composite (PVDF/HTO) have been investigated by using differential scanning calorimetric, wide-angle X-ray diffraction, light scattering, and infrared spectroscopic analyses. The neat PVDF predominantly formed α-phase in the crystallization temperature range of 110-150 °C. On the other hand, PVDF/HTO exhibited mainly α-phase crystal coexisting with γ- and β-phases at low Tc range (110-135 °C). A major γ-phase crystal coexists with β- and α-phases appeared at high Tc (=140-150 °C), owing to the dispersed layer titanate particles as a nucleating agent. The overall crystallization rate and crystalline structure of pure PVDF were strongly influenced in the presence of layered titanate particles.  相似文献   

6.
D. Kilburn  G. Dlubek  J. Pionteck 《Polymer》2006,47(22):7774-7785
Free volume data from positron annihilation lifetime spectroscopy (PALS) experiments are combined with a Simha-Somcynsky (S-S) equation of state analysis of pressure-volume-temperature (PVT) data to model free volume contributions to structural mobility in a series of poly(n-alkyl methacrylate)s. From the PALS data the glass transition temperature, Tg, decreases (from 382 to 224 ± 5 K) and a given mean free volume is observed at lower temperatures as the side-chain length increases (going from methyl- to hexyl-). This is evidence of an internal plasticization whereby the side-chains reduce effective packing of molecules. By comparing PALS and PVT data, the hole number per mass unit, Nh′, is calculated using different methods; this varies between 0.54 and 0.86 × 1021 g−1. It is found that the extrapolated free volume becomes zero at a temperature T0′ that is smaller than the Vogel temperature T0 of the α-relaxation. The α-relaxation frequencies can be fitted by the free volume theory of Cohen and Turnbull, but only when the free volume Vf is replaced by (Vf − ΔV) where ΔV( = Ef(T0 − T0′), Ef is the thermal expansivity of Vf) varies between 0.060 and 0.027 ± 0.003 cm3/g, decreasing with side-chain length, apart from poly(n-hexyl methacrylate) where ΔV increases to 0.043 ± 0.003 cm3/g. One possible interpretation of this is that the α-relaxation only occurs when, due to statistical reasons, a group of m or more unoccupied S-S cells are located adjacent to one another. m is found to vary between 8 and 2 for poly(methyl methacrylate) and poly(n-butyl methacrylate), respectively. We found that no specific feature in the free volume expansion was consistently in coincidence with the dynamic crossover.  相似文献   

7.
A calorimetric, dielectric and dynamic-mechanical study of the dynamics of the poly(t-butyl acrylate) (PtBa) chains has been carried out in a PtBa homopolymer and two polystyrene (PS)-b-PtBa block copolymers with different PtBa chain lengths. The DSC results show that the size of the cooperative rearranging regions is similar in the homopolymers and the copolymers, both for the PtBa rich- and the PS-rich regions. Therefore, no significant contributions are found arising from composition fluctuations in the copolymers. The relaxation map obtained from dielectric relaxation indicates that there are no differences in the temperature dependence of the α-relaxation of the PtBa block in the three samples studied. However, there are larger differences for the values obtained from DMTA experiments. Contrary to the α-relaxation, the relaxation map for the β-transition shows that the characteristic times for the PtBa blocks are smaller in the homopolymer than in the copolymers. In principle, these are unexpected results because the β-relaxations have a more local character than the α-ones. The width of the α-relaxation increases with T for all the samples, and it is slightly larger for the copolymers. The intensity of the α-relaxation is larger (between 3 and 4 times) for the homopolymer. Considering the molecular weights of the PtBa blocks, this effect has to be ascribed to the existence of frozen amorphous PtBa due to the existence of the glassy PS domains in the microphase separated copolymers.Molecular Dynamic Simulations (MDSs) for different sequences of the polymers under study were carried out. The conformational analysis was carried out between 1000 and 1700 K. The analysis of the variation of angles ?1 and ?2 of the ester group of PtBa points out the existence of a correlation between the conformational changes of the side group of the polymer chains and their relaxational behaviour.  相似文献   

8.
Florian J. Stadler 《Polymer》2005,46(23):10311-10320
Polyethylenes and polyethylene/α-olefin-copolymers covering a range in crystallinity between 12 and 85% were investigated by means of dynamic-mechanical measurements between −145 °C and their melting point. From the temperature and frequency dependence of the complex modulus α′-, α-, β- and γ-relaxations were analyzed. The α′-relaxation was discovered in all HDPE-, LDPE- and LLDPE-samples but not in plastomer- and elastomer-samples. The activation energies (30-140 kJ/mol) of this relaxation were found to decrease with increasing crystallinity. The α′-transition temperature at a fixed frequency rises with increasing degree of crystallinity and tends to reach the melting point when approaching the fully crystalline state. Thus, it is concluded that the α′-relaxation originates from the interface between crystal lamellae and amorphous interlamellar regions. By extrapolation of the storage modulus to the amorphous state the entanglement molar mass was calculated as 2300 g/mol for a completely amorphous polyethylene/α-olefin-copolymer.  相似文献   

9.
João F Mano  João L Lopes 《Polymer》2003,44(15):4293-4300
Isothermal short-term creep of poly (vinylidene fluoride) (PVDF) monofilament sutures was determined at several temperatures between 10 and 90 °C under the stress of 10 MPa. Long term service performance was predicted for 10 decades of time. The compliance master curve as a function of time fits a hyperbolic sine equation. The temperature shift factor as a function of the temperature aT (T) is accurately represented by a general equation based on free volume. A simple relationship between the two parameters of the equation is explored. The viscoelasticity of PVDF is also seen in dynamic mechanical analysis performed at the frequency of 1 Hz. The origin of the viscoelastic character well present in the deformability of the PVDF in service is due to the occurrence of the αc relaxation that is active at ∼50 °C (E″ peak at 1 Hz).  相似文献   

10.
Kai C. Yen 《Polymer》2009,50(2):662-98
Polymorphism and its influential factors in poly(heptamethylene terephthalate) (PHepT) were probed using differential scanning calorimetry (DSC), Fourier-transform infrared (FTIR) spectroscopy, and wide angle X-ray diffraction (WAXD). PHepT exhibits two crystal types (α and β) upon crystallization at various isothermal melt-crystallization temperatures (Tcs) by quenching from different Tmaxs (maximum temperature above Tm for melting the original crystals). Melt-crystallized PHepT with either initial α- or β-crystal by quenching from Tmax lower than 110 °C leads to higher fractions of α-crystal, but crystallization from Tmax higher than 140 °C leads to higher fractions of β-crystal. In addition to Tmax, polymorphism in PHepT is also influenced by crystallization temperature (Tc = 25-75 °C). When PHepT is melt-crystallized from a high Tmax = 150 °C (completely isotropic melt), it shows solely β crystal for higher Tc, and solely the α-crystal for Tc < 25 °C; in-between Tc = 25 and 35 °C, mixed fractions of both α- and β-crystals. However, by contrast, when PHepT is melt-crystallized from a lower Tmax = 110 °C, it shows α-crystal only at all Tcs, high or low.  相似文献   

11.
The preparation of polyimides containing side-chain chromophores and the long-term aging performance of poled films are described. These materials were compared to guest-host polycarbonate films. Mach-Zehnder optical interferometers were fabricated from these polymers that contained CLD- and FTC-type chromophores. Changes in optical properties were monitored for months at four temperatures ranging from ambient to 110 °C. The isothermal relaxation data were modeled using both a stretched exponential equation and a power law in time equation. The temperature dependency of the time constants of these equations was modeled using a new activation-energy equation: ln(τ/τp) = ER(1 + tanh[(Tc − T)/D])/2RT + Ep/RT where Tc is the central temperature of the transition zone, D is the breadth of the zone, and Es are the activation energies of rigid and pliable materials. Multi-year high-temperature stability of the poled guest-host and side-chain materials was predicted.  相似文献   

12.
The structures of (OCH2CHR) with R=CH2S(CH2)6SO2(CH2)MH (ATP-M; M=5,7,9) or R=CH2SO2(CH2)6SO2(CH2)MH (ASP-M; M=5,7,9) were studied using X-ray diffraction and differential scanning calorimetry. The X-ray patterns of all ATPs and ASPs studied show a series of ordered reflections in the small angle region and a sharp wide angle reflection at d=∼4.4 Å, characteristic of a smectic phase. The smectic layer thickness corresponds to twice the most extended side chain length and linearly increases as the side chain length increases with a slope of ∼2.3 Å per methylene spacer. This indicates that all ASPs and ATPs studied have a double layer structure with side chains normal to the main chain and probably an all-trans conformation of the side chains. The correlation lengths measured from the wide angle reflections are in the range of 80±10 Å for all the polymers except for ASP-5 (∼40 Å). These values indicate that quasi-long-range order exists in the smectic layers whose structures can be defined as smectic B (SB). The d-spacing of the wide angle reflection, 4.4 Å, suggests that paraffinic side chain crystallization does not occur and that the smectic mesophase develops through dipole-dipole interactions of sulfone groups in the side chains. During heating, ATP-5 shows recrystallization after the first melting. The structure produced during recrystallization has a similar smectic structure but with more dense packing between side chains than before the first melting. In the case of ASP-9, a smectic to smectic transition was observed at ∼110°C prior to the isotropic temperature at ∼150°C. Both the correlation length (from the wide angle reflection) and the layer thickness decreased from ∼80 to ∼30 Å and from 46 to 40 Å at this transition, respectively, indicating that the order in the smectic layers is lost and the SB structure has become a less ordered SA structure at this transition.  相似文献   

13.
Lead-free (Ba0.85Ca0.15)(Ti1−xZrx)O3 (BCTZ) piezoelectric ceramics were fabricated by normal sintering in air atmosphere. BCTZ ceramics with x = 0.10 possess a coexistence of tetragonal and rhombohedral phases at ∼40 °C. The Curie temperature of BCTZ ceramics decreases with increasing the Zr content. Piezoelectric properties of BCTZ ceramics are dependent on the poling conditions (i.e., the poling temperature and the poling electric field), and the underlying physical mechanism is illuminated by the phase angle. The BCTZ (x = 0.10) ceramic, which locates at the existence of two phases and is poled at E ∼ 4.0 kV/mm and Tp ∼ 40 °C, exhibits an optimum electrical behavior at a room temperature of ∼20 °C: d33 ∼ 423 pC/N, kp ∼ 51.2%, 2Pr ∼ 18.86 μC/cm2, 2Ec ∼ 0.47 kV/mm, ?r ∼ 2892, and tan δ ∼ 1.53%.  相似文献   

14.
Samy A. Madbouly 《Polymer》2007,48(14):4097-4107
The miscibility and molecular dynamics of nanostructured maleated polypropylene (mPP)/polycarbonate (PC) blends prepared by in situ polymerization of macrocyclic carbonates with polypropylene modified with 0.5 wt% of maleic anhydride-reactive groups were investigated over a wide range of frequencies (10−2-0.5 × 107 Hz) at different constant temperatures using broadband dielectric spectroscopy and scanning transmission electron microscope (STEM). The molecular dynamics of the glass relaxation process of the blend (α-relaxation process) appeared at a lower temperature range compared with that of the pure PC. This shift in the molecular relaxation process is attributed to the partial miscibility of the two polymer components in the blends as previously confirmed by the morphology via STEM. Nanoscale morphologies with average domain diameters as small as 50 nm were obtained for the different blend compositions studied. The STEM photographs show that the graft mPP-g-PC prefers to locate at the interfaces as previously reported. The relaxation spectrum of pure PC and mPP/PC blends was resolved into α- and β-relaxation processes using the Havriliak-Negami equation and ionic conductivity. The dielectric relaxation parameters, such as relaxation peak broadness, maximum frequency, fmax, and dielectric strength, Δ? (for the α- and β-relaxation processes), were found to be blend composition dependent. The kinetics of the α-relaxation processes of the blends were well described by Vogel-Fulcher-Tammann (VFT) equation. The local process of PC was resolved into two relaxation processes β1 and β2, associated with the carbonyl groups' motion and the combined motions of carbonyl and phenylene groups, respectively. Only β2 shifted to lower frequency in the blend while β1 was relatively not affected by blending. The electric modulus of the blends was used to get a sufficient resolution of the different relaxation processes in the samples, i.e., α-, β-relaxation processes, ionic conductivity, and interfacial polarization. In addition, the blending method used was found to increase the d.c. conductivity without affecting the charge carrier transport mechanism, making it possible to develop novel polymer blends with tunable dielectric properties and morphology from existing polymers.  相似文献   

15.
The structure of a side chain liquid crystalline poly(silylenemethylene) (-(SiCH3R-CH2)-: R=O(CH2)11O-Ph-Ph-CN, Ph=phenyl) (CN-11) has been studied by X-ray diffraction and differential scanning calorimetry (DSC). The DSC results showed that CN-11 has transitions at ∼92 °C (T2) and ∼147 °C (T1) during both cooling and immediate heating. A third transition occurred at ∼50 °C (T3) during heating after annealing at room temperature. The X-ray fiber pattern of the CN-11 annealed at room temperature showed several wide and small angle reflections which were indexed by a monoclinic unit cell with parameters a=16.8 Å, b=7.42 Å, c=43.6 Å and β=102.1° (b: fiber direction), representing a crystal structure with layer thickness of ∼43 Å. Upon heating at T3, the crystal structure became less ordered (but somewhat more ordered than smectic A (SA) and smectic C (SC)). This was followed by SA (or SC) phase at T2, and ultimately an isotropic state (I) at T1. The observed layer thickness (∼43 Å) is about ∼1.5 times the most extended side chain length, indicating a double-layer structure with tilted or interdigitated side chains. The X-ray fiber pattern had a four-point pattern at d=4.52 Å, suggesting that the side chains in the crystal are likely to be tilted by 56° from the polymer fiber axis.  相似文献   

16.
Impedance spectroscopy measurements were carried out in the 10 Hz to 10 MHz frequency range from 500 to 1200 °C in LaAl11O18 pellets sintered at 1600 °C. The powders were obtained by the polymeric precursor technique. The sintered pellets were nearly single phase LaAl11O18. The bulk electrical resistivity was evaluated from the [−Z″(ω) × Z′(ω)] impedance diagrams. The value of the activation energy for the ionic conduction, 0.89 eV, was determined from the Arrhenius plot of the bulk conductivity. An yttria-stabilized zirconia (YSZ) oxygen pump and an YSZ oxygen sensor were used for providing 10–1500 ppm of partial pressure of oxygen (pO2) at 1000 °C for determining the electromotive force (emf) in a Pt/LaAl11O18/Cr2O3 + Cr electrochemical cell. The results follow the Nernst law. The high signal-to-noise ratio of the emf at low pO2 values shows that the LaAl11O18 specimens may be used in sensors for detection of oxygen at high temperatures.  相似文献   

17.
Water vapor oxidation and salt corrosion resistances of mullite-gadolinium silicate (Gd2SiO5) environmental barrier coatings (EBCs) dip coated on α-SiC substrates and sintered to 1430 °C/3 h in air were investigated. The EBC exhibited excellent adherence to the substrate during thermal cycling between 1350 °C and room temperature (RT) for 100 h in a simulated lean combustion environment (90% H2O-balance O2), [11], [13] and [15] forming ∼10 μm porous silica layer at coating-substrate interface, compared to ∼17 μm for uncoated α-SiC exposed under same conditions. The EBC did not spall after a 24 h 1200 °C exposure in Na2SO4 corrosion environment leading to further coat densification. However, after 48 h salt exposure, the EBC showed severe through-thickness cracks, and cavities and de-lamination at coating-substrate interface. The corrosion gaseous products such as CO2, CO and SO2 trapped under a low viscosity glassy (Na2x(SiO2)) liquid phase were formed due to salt vapor reaction with α-SiC substrate created these cavities.  相似文献   

18.
K.L. Ngai  T.R. Gopalakrishnan 《Polymer》2006,47(20):7222-7230
The voluminous amount of data in the literature on the structural α- and the Johari-Goldstein β-relaxations of the poly(n-alkyl methacrylate)s allows a systematic study of the interrelation between the two important relaxation processes. The data bring out the systematic changes in the interrelation between the structural α- and the Johari-Goldstein β-relaxations with changes in molecular structure, molecular weight, tacticity and size (by nanoconfinement), and modifications by copolymerization, and crosslinking. The results can all be interpreted as primarily due to changes in intermolecular coupling, which have significant effects on the many-molecule dynamics constituting the structural α-relaxation, but not on the precursory Johari-Goldstein β-relaxation. Theoretically, the Coupling Model predicts a relation of intermolecular coupling (or degree of cooperativity of the α-relaxation) to the ratio of the α- and the β-relaxation times, and a correlation of intermolecular coupling to the steepness or “fragility” index. The predicted relation and correlation are compared with experimental data of the poly(alkyl methacrylate)s.  相似文献   

19.
Barium titanate (BaTiO3/BT) ferroelectric system was synthesized in single perovskite phase at low temperature by using powders derived from modified solid state reaction (MSSR) and sintered by microwave (MW) processing routes. Conventional calcination temperature was optimized at 900 °C for 4 h. MW sintering of BT samples was carried out at 1100 °C for 30 min to get dense (98% density) ceramics. Room temperature (RT) dielectric constant (?r) and dielectric loss (tan δ) at 1 kHz frequency of MW sintered BT samples was found to be ∼2500 and 0.03, respectively. Saturated polarization vs. electric field (P-E) loops with remnant polarization (Pr) ∼6 μC/cm2 and coercive field (Ec) ∼1.45 kV/cm confirmed the ferroelectric nature of MW sintered BT samples. Piezoelectric coefficient from strain vs. electric field (S-E) loops study was found to be 335 pm/V.  相似文献   

20.
The effects of incorporated poly(d-lactic acid) (PDLA) as poly(lactic acid) (PLA) stereocomplex crystallites on the isothermal and non-isothermal crystallization behavior of poly(l-lactic acid) (PLLA) from the melt were investigated for a wide PDLA contents from 0.1 to 10 wt%. In isothermal crystallization from the melt, the radius growth rate of PLLA spherulites (crystallization temperature (Tc)≥125 °C), the induction period for PLLA spherulite formation (ti) (Tc≥125 °C), the growth mechanism of PLLA crystallites (90 °C≤Tc≤150 °C), and the mechanical properties of the PLLA films were not affected by the incorporation of PDLA or the presence of stereocomplex crystallites as a nucleating agent. In contrast, the presence of stereocomplex crystallites significantly increased the number of PLLA spherulites per unit area or volume. In isothermal crystallization from the melt, at PDLA content of 10 wt%, the starting, half, and ending times for overall PLLA crystallization (tc(S), tc(1/2), and tc(E), respectively) were much shorter than those at PDLA content of 0 wt%, due to the increased number of PLLA spherulites. Reversely, at PDLA content of 0.1 wt%, the tc(S), tc(1/2), and tc(E) were longer than or similar to those at PDLA content of 0 wt%, probably due to the long ti and the decreased number of spherulites. This seems to have been caused by free PDLA chains, which did not form stereocomplex crystallites. On the other hand, at PDLA contents of 0.3-3 wt%, the tc(S), tc(1/2), and tc(E) were shorter than or similar to those at PDLA content of 0 wt% for the Tc range below 95 °C and above 125 °C, whereas this inclination was reversed for the Tc range of 100-120 °C. In the non-isothermal crystallization of as-cast or amorphous-made PLLA films during cooling from the melt, the addition of PDLA above 1 wt% was effective to accelerate overall PLLA crystallization. The X-ray diffractometry could trace the formation of stereocomplex crystallites in the melt-quenched PLLA films at PDLA contents above 1 wt%. This study revealed that the addition of small amounts of PDLA is effective to accelerate overall PLLA crystallization when the PDLA content and crystallization conditions are scrupulously selected.  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