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 objective of this work was to develop biodegradable films using fibrous residue from the extraction of cassava starch and to investigate its effect as a reinforcement agent in polymeric films. Micrographs (SEM) showed that the obtained films presented smooth surfaces with no cracking. The addition of fibers affected the properties of the films significantly (p < 0.05), reducing water vapor permeability values (from 8.63 ± 0.15 × 10−11 g.m.m−2.s−1.Pa−1 to 3.33 ± 0.16 × 10−11 g.m.m−2.s−1.Pa−1) and increasing the maximum tensile strength (from 1.23 ± 0.15 MPa to 7.78 ± 0.83 MPa). These results encourage the use of cassava bagasse as a reinforcement in the production of green composites as packaging. © 2018 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2019 , 136, 47224.  相似文献   

2.
Poly(L ‐lactide) (PLLA) films having different crystallinities (Xc's) and crystalline thicknesses (Lc's) were prepared by annealing at different temperatures (Ta's) from the melt and their high‐temperature hydrolysis was investigated at 97°C in phosphate‐buffered solution. The changes in remaining weight, molecular weight distribution, and surface morphology of the PLLA films during hydrolysis revealed that their hydrolysis at the high temperature in phosphate‐buffered solution proceeds homogeneously along the film cross‐section mainly via the bulk erosion mechanism and that the hydrolysis takes place predominantly and randomly at the chains in the amorphous region. The remaining weight was higher for the PLLA films having high initial Xc when compared at the same hydrolysis time above 30 h. However, the difference in the hydrolysis rate between the initially amorphous and crystallized PLLA films at 97°C was smaller than that at 37°C, due to rapid crystallization of the initially amorphous PLLA film by exposure to crystallizable high temperature in phosphate‐buffered solution. The hydrolysis constant (k) values of the films at 97°C for the period of 0–8 h, 0.059–0.085 h–1 (1.4–2.0 d–1), were three orders of magnitude higher than those at 37°C for the period of 0–12 months, 2.2–3.4×10–3 d–1. The melting temperature (Tm) and Xc of the PLLA films decreased and increased, respectively, monotonously with hydrolysis time, excluding the initial increase in Tm for the PLLA films prepared at Ta = 100, 120, and 140°C in the first 8, 16, and 16 h, respectively. A specific peak that appeared at a low molecular weight around 1×104 in the GPC spectra was ascribed to the component of one fold in the crystalline region. The relationship between Tm and Lc was found to be Tm (K) = 467·[1–1.61/Lc (nm)] for the PLLA films hydrolyzed at 97°C for 40 h.  相似文献   

3.
Steady state fluorescence (SSF) and UV–visible techniques have been used to study neck growth and equilibration processes during the coalescence of hard latex particles. Latex films were prepared separately by annealing pyrene (Py) labelled and unlabelled poly(methyl methacrylate) (PMMA) particles above their glass transition temperature. During the annealing processes, the optical clarity of the films increased considerably. Direct fluorescence emission of excited pyrene from labelled latex films was monitored as a function of annealing temperature to detect this change. Void closure temperature (Tc) and time (tc) were determined at the point where the fluorescence emission intensity became maximal. Below this point, the increase in fluorescence intensity (Iop) against temperature was used to determine the activation energy for viscous flow (ΔH≈47kcalmol−1). The decrease in Iop above the void closure temperature was used to determine the backbone activation energy (ΔE≈44kcalmol−1) for the interdiffusing chains. Unlabelled PMMA particles were used to prepare films for UV–vis measurements. The transmitted photon intensity (Itr) from these films increased as the annealing temperature was increased. This behaviour was also used to determine the backbone activation energy (ΔE≈35kcalmol−1) for the interdiffusing chains. © Society of Chemical Industry.  相似文献   

4.
The absorption and subsequent desorption behaviors of amorphous polymer films of PEEK poly(ether ether ketone), PEEKK poly(ether ether ketone ketone), and PEKK poly(ether ketone ketone) in solvent of 1,2-dichloroethane (C2H4Cl2) are investigated and compared. The equilibrium absorption weight (M) of these polymers is related to their molecular ketone content or molecular chain rigidity and also to the experimental conditions. Especially, at a certain temperature, the molecular chains in the solvent can be polarized, which leads to producing greater M for polymer films; for example, at 60°C, M = 46% for PEEK and M = 65% for PEKK. The pseudodiffusion coefficients for PEEK, PEEKK, and PEKK all surpass the 6.0 × 10−12 m2 s−1. The polymer's molecular polarization has been proved in concentrated sulfur acid. Results also show that amorphous resin's films become white and creeped in dichloroethane, which is more serious when metaphenyl links are introduced into PEEKK or PEKK molecular main chains. The residual solvent of 1% or so often exists in the films, even though a long desorption time (over 100 h) has been proceeded. Absorption has induced crystallization of amorphous polymer films, but this crystallization process is slightly different from that of the films crystallized from both the glassy state and the melting state in the solvent, which makes the amorphous interlayers grow progressively and more condensely; thus, the crystallized films will have higher Tg's than these crystallized under annealing condition. The morphology results have shown that the solvent-crystallized films are less toughened than the amorphous ones because of the intermediate layer between the induced crystallized area and the amorphous area in the core. © 1998 John Wiley & Sons, Inc. J Appl Polym Sci 67:2065–2075, 1998  相似文献   

5.
UV–visible (UVV) technique is used for monitoring of polymer film dissolution. These films are formed from pyrene (P)‐labeled poly(methyl methacrylate) (PMMA) latex particles, sterically stabilized by poly isobutylene (PIB, Annealing of films was performed above Tg at various temperatures for 30‐min time intervals. Diffusion of solvent molecules (chloroform) into the annealed latex film was followed by desorption of PMMA chains. Desorption of P‐labeled PMMA chains was monitored in real time by the absorbance change of P in the polymer–solvent mixture. A diffusion model with a moving boundary was used to quantify real time UVV data. Diffusion coefficients of desorbed PMMA chains were measured and found to be between 2 and 0.6 × 10−11 cm2 s−1 in the 100 and 275°C temperature range. © 2000 John Wiley & Sons, Inc. J Appl Polym Sci 75: 1075–1082, 2000  相似文献   

6.
This article reports on using cassava starch nanocrystals (CSN) to strengthen nanocomposite films from the same matrix. CSN were obtained by acid hydrolysis. Nanocomposite (starch:glycerol:CSN/4.0:2.1:1–10 wt %) were processed by casting and the films were characterized. The CSN (30% yield) presented minimally clustered globular forms, 45 to 178 nm in diameter, with a crystalline index of 46%. Water‐vapor transmission rate, tensile strength, and elastic modulus of the films were influenced by the linear effect of CSN concentration (R 2 = ?0.92, 0.91, 0.99, respectively), while the other parameters resulted in quadratic relations |0.69–0.96|. The film with 10% CSN presented a 43% reduction in water vapor permeability, associated with increases of 200% in traction resistance, and 616% in elasticity modulus, compared with the control. The hydrolysis of part of the cassava starch into nanocrystals resulted in a reduction in permeability and nano reinforcement of the films due to good compatibility and interaction between both, without influencing biodegradability. © 2017 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2017 , 134 , 45311.  相似文献   

7.
Tensile mechanical properties of poly(aryl ether ether ketone) (PEEK) films showing different thermal histories have been investigated at room temperature to point out the main key microstructural features governing properties over a wide strain rate range, i.e., from 10−5 to 300 s−1. The strain rate sensitivity of the mechanical properties of amorphous PEEK films significantly depends on the analyzed strain rate range: i.e., 1) from 10−5 to 10 s−1, the strain rate dependence of both apparent Young's modulus and yield stress is weak; and 2) from 10−1 to 200 s−1, both parameters significantly increase. Thus, based on the definition of the relationships between temperature, strain rate, and frequency respectively used for tensile tests and dynamic mechanical spectrometry, it was shown that the mechanical behavior of PEEK films at room temperature could be governed by similar molecular mechanisms as those giving rise to the β1 and β2 transitions. The Eyring analysis shows that motions of five or six monomers are implied at the beginning of the plastic deformation of amorphous and semi-crystalline PEEK films, while at higher strain rates, shorter chain segments are concerned. Thus, the crystalline phase only induces an increase in the stress level because of the reinforcement effect but does not modify the molecular mechanisms governing the plastic deformation of PEEK films at room temperature. © 1997 John Wiley & Sons, Inc. J Appl Polym Sci 64: 1053–1059, 1997  相似文献   

8.
Films based on bitter potato starch (BPS) and its blends with chitosan (BPS-Ch) or soy protein isolate (BPS-SPI) loaded with copaiba leaf extract (E) are prepared via the casting method. The physical–mechanical and antioxidant properties of the as-prepared films are compared with those of a control. Moreover, the half-maximal degradation (t50) of the prepared films is calculated by fitting the Hill model to disintegrability kinetic data. Among the analyzed films, BPS-Ch-E exhibits the lowest (p < 0.05) solubility in water and opacity, strongest water vapor-barrier (3.58 × 10−11 g m−1 s−1 Pa−1), and highest tensile strength and elongation at break. The Fourier transform infrared spectra of BPS-Ch-E and BPS-SPI-E demonstrate new peaks at 1550, 1239, and 1070 cm−1 corresponding to N H and C O stretching. The BPS-E and BPS-Ch-E surfaces are devoid of scratches and phase separation. The incorporation of E significantly increases the antioxidant activity of the films. BPS-SPI-E and BPS-Ch-E demonstrate the lowest (t50 ≈ 1.4 days) and highest (t50 ≈ 3.5 days) disintegration rates, respectively, among the prepared films. E loading facilitates the development of films possessing beneficial physical–mechanical and antioxidant properties as well as rapid disintegrability, enabling their potential application as a eco-friendly packaging material.  相似文献   

9.
The effects of crystallinity (xc) on the hydrolysis of high molecular weight poly(L ‐lactide) (PLLA) films in a phosphate‐buffered solution at 37°C was investigated by gel permeation chromatography, tensile testing, differential scanning calorimetry, scanning electron microscopy, and polarizing optical microscopy. The change in molecular weight distribution and surface morphology of the PLLA films after hydrolysis revealed that the hydrolysis of PLLA film in a phosphate‐buffered solution proceeded homogeneously along the film cross section, mainly via the bulk‐erosion mechanism. The induction period until the start of the decrease in mass remaining and the tensile strength became longer with a decrease in the initial xc of the PLLA films. The rate of molecular weight reduction was higher as the initial xc of the PLLA films increased when hydrolysis was carried out up to 24 months. Melting and glass transition temperatures of the PLLA films increased in the first 12 months of hydrolysis, while they decreased in another 24 months, irrespective of the initial xc. The xc value of the PLLA films increased monotonously by hydrolysis. The lamella disorientation in PLLA spherulites after hydrolysis implied that the hydrolysis of PLLA chains occurred predominantly in the amorphous region between the crystalline regions in the spherulites. The area of a specific molecular weight in GPC spectra at 36 months increased with increase in the initial xc of the PLLA film, suggesting that the specific peak should be due to the component of one fold in the crystalline region. The reason for enhanced hydrolysis of PLLA films having higher initial crystallinities was discussed in terms of tie chains and terminal groups of PLLA. © 2000 John Wiley & Sons, Inc. J Appl Polym Sci 77: 1452–1464, 2000  相似文献   

10.
The phase behaviour of two classes of side chain polysiloxanes, containing respectively 4-cyano- and 4-alkoxy-4′-stilbene-ether mesogens, was studied over a wide temperature range by differential scanning calorimetry and wide angle X-ray diffraction, with special attention devoted to the characterisation of the microdomain phase morphology. Room temperature X-ray diffraction studies suggest for certain polymers with cyanostilbene mesogens the occurrence of microphase separation; this effect becomes significant as the amount of polysiloxane backbone versus the side chains increases. The onset of electrical conductivity was followed upon exposure of the polymeric films to various doping agents; conductivities up to 2 × 10−1 S cm−1 were measured when cyanostilbene-containing polymers were exposed to strong Lewis acids such as SO3 or SbF5. However, occurrence of some chemical degradation upon heavy doping and of electrode polarisation during DC conductivity measurements indicate that the observed electroconductivity is not purely electronic.  相似文献   

11.
The permeability coefficients of O2, N2, and CO2 gases at 25°C were examined for composite membranes that were prepared by filling poly(ethylene oxide)(PEO) with different molecular weights into a porous membrane. The permeability coefficients of O2, N2, and CO2 were 2 × 10−10 – 4 × 10−10, 5 × 10−11 – 9.5 × 10−11, and 6 × 10−10 – 1 × 10−9 (cm3 STPcm/cm2 s cmHg), respectively. The higher permeability coefficients of CO2 are explained in terms of high solubility of CO2 in filled PEO. The permeability coefficient of CO2 was affected by the degree of crystallinity of PEO in the composite. On the other hand, there was little effect of crystallinity on O2 and N2 permeability coefficients. Some probable relationships between selectivities of O2 to N2 and CO2 to N2 and the degree of crystallinity of PEO were observed. The CO2 gas permeability coefficients of the composite membrane for PEO50000 (Mw = 5 × 104) showed a marked change due to melting or crystallization of PEO. © 1999 John Wiley & Sons, Inc. J Appl Polym Sci 73: 2733–2738, 1999  相似文献   

12.
The wide molar mass distribution of native starch creates obstacles in investigating the physicochemical characteristics of starch, such as retrogradation, because samples thought to be the same are actually compounds containing many chains with different molar masses. In this paper, the sweet potato amylose and amylopectin isolated from retrograded starch were treated with the retrogradation–hydrolysis method three times, and their physicochemical changes in this process were determined by absorbance of the starch–iodine complex, light microscopy, and molar mass and chain length distributions. The results showed that repeated retrogradation and hydrolysis caused the molar mass distribution of sweet potato amylose and amylopectin to reduce from 4.2 × 107–205 and 7971–223 to 6.0 × 104–730 and 4533–211 g mol−1, respectively. This retrogradation–hydrolysis cut the chain length distribution of sweet potato amylose from DP 9–35 to DP 3–13, but that of amylopectin remained unchanged. The double helix in sweet potato amylopectin will not form if the percentage of chain length with DP ≥ 4 is less than 25%. Repeated retrogradation and hydrolysis was an appropriate method to obtain amylose or amylopectin with a narrower molar mass distribution. © 2016 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2016 , 133, 43849.  相似文献   

13.
Dielectric and calorimetric measurements have been carried out for tetramethyl polycarbonate/polystyrene (TMPC/PS) blends with different compositions. The effect of varying the molecular weight of the weakly polar component (PS) on the molecular dynamics of the polar segments of TMPC has been thoroughly studied over wide ranges of frequency (10−2−105 Hz), temperature (50–220°C) and number average molecular weight, n, (6500–560 000 g mol−1). All blends were found to be compatible regardless of the composition ratio and the molecular weight of PS. Some new and interesting experimental findings have been observed concerning the effect of molecular weight on the glass temperature and on the broadness of the glass transition and relaxation. Neither the kinetics nor the distribution of relaxation times of the local process observed in pure TMPC was affected by blending with PS, regardless of the composition ratio or the molecular weight of PS. It has been concluded that the mixing of the polymeric components to form a homogeneous single phase (compatible blend) does not take place on a segmental level but on a structural one. The size of this structural level has been suggested to have the same volume as the cooperative dipoles, which is assumed to be the minimum volume responsible for a uniform glass transition (10–15 nm). The molecular weight dependence of the relaxation characteristics of the glass process and temperature could be attributed to the variation in the size and packing of the structural units.  相似文献   

14.
A new technique, based on steady-state fluorescence measurements, is introduced for studying dissolution of polymer films. These films are formed from naphthalene and pyrene labeled poly(methyl methacrylate) (PMMA) latex particles, sterically stabilized by polyisobutylene. Diffusion of solvent (chloroform) into the annealed latex film was followed by desorption of polymer chains. Annealing was performed above Tg at various temperatures for 30-min time intervals. Desorption of pyrene labeled PMMA chains was monitored in real time by the pyrene fluorescence intensity change. Desorption coefficients were found to be between 1 and 4 × 10−10 cm2/s and two different dissolution mechanisms were detected. © 1996 John Wiley & Sons, Inc.  相似文献   

15.
Graft copolymerization of 4-vinyl pyridine (4-VP) on poly(ethylene terephthalate) (PET) films using a benzoyl peroxide (Bz2O2) initiator was investigated under different conditions including polymerization time, temperature, monomer concentration, and initiator concentration. Dimethyl sulfoxide (DMSO) was used as swelling agent to promote the incorporation and the subsequent polymerization of 4-VP to PET films. Maximum percent grafting was obtained when the polymerization was carried for a period of two hours at 65°C. Increasing the monomer concentration from 0.2M to 0.8M and Bz2O2 concentration from 1.0×10−3M to 2.5×10−3M was accompanied by a significant enhancement in percent grafting. Monomer diffusion on PET films and its effect on the grafting yield were studied and intrinsic viscosities of grafted films were also measured. © 1996 John Wiley & Sons, Inc.  相似文献   

16.
To endow waterborne fluorinated polyurethanes (WFPUs), films with stable lower surface free energy and lower surface reorganization after the films contacted with water, a series of crosslinked WFPUs (CWFPUs) emulsions were prepared by adjusting the content of aziridine (AZ). The effect of crosslinking on the surface free energy, glass-transition temperature (T g), water absorption, and surface composition of dry/hydrated WFPUs and CWFPUs films were studied by CAs test, dynamic mechanical analysis, water absorption measurements, and X-ray photoelectron spectroscopy. When the fluorine content was 0.5%, the surface free energy of the CWFPUs films modified by 0.4% AZ content (CWFPU-6-0.5-CK0.4) reached to the lowest value of 15.76 mN m−1 which almost equaled to the surface free energy (15.45 mN m−1) of the dry uncrosslinked films (WFPU-6-0.5). With the increasing of AZ content, the T g, hard of hard segments of the CWFPUs films increased and the water absorption of the CWFPUs films decreased, which suggested the formation of the crosslinked network structures. The studies of the surface elements and groups composition of dry/hydrated WFPUs and CWFPUs both confirmed that the surface reorganization and the migration of fluorinated side chains were restricted by the crosslinked network structures. © 2018 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2019 , 136, 47167.  相似文献   

17.
In order to meet the requirements of polymer dielectric materials for high thermal stability and excellent dielectric properties in the application of high-temperature film capacitors, a series of polyimide (PI) films are fabricated by introducing a self-synthesized aniline trimer (ACAT) with a conjugated structure in this work. Since the conjugated ACAT in the main chains of PI improves the electron polarization and carrier mobility of the PI molecular chains, the dielectric constant of the ACAT-PI films is greatly enhanced (4.4–7.4). Meanwhile, the dissipation factor does not increase apparently (0.002–0.013). The dielectric properties are stable even when the temperature is up to 200 °C, the thermal degradation temperature is as high as 450 °C, and the mechanical properties are also excellent (70–105 MPa). Among all the films, the PI film with 5 mol% ACAT exhibits the maximal energy density of 3.6 J cm−3 under the field of 426 kV mm−1, the high tensile strength (90 MPa) and the excellent thermal stability (Td5 = 515 °C). The work paves the way to prepare high-temperature polymer dielectric film materials with high energy storage density.  相似文献   

18.
Carbon molecular sieve membranes (CMSM) were prepared from the pyrolysis of polyimide films within a temperature range of 600°C-800°C under nitrogen stream. The membrane samples were characterized and tested for the permeation of He, CH4, CO2, and N2 at different pressures and temperatures, respectively. The CMSM700 membrane (pyrolyzed at 700°C) showed an ideal selectivity of ~ 11 for N2/CH4 with a permeability of 2.18 × 10−15 mol · m/m2 · s · Pa for N2. The separation mechanism for the N2/CH4 pair was shown to be largely molecular sieving rather than surface flow. The membrane showed an ideal selectivity of ~ 500 for the CO2/CH4 pair with a CO2 permeability of 9.72 × 10−14 mol · m/m2 · s · Pa. The permeability of He was lower than that of CO2, suggesting that the surface flow played a significant role in the CO2 permeation. The updated permeability-selectivity tradeoff curves show that this CMSM membrane compared favourably with other membrane materials reported in the literature for the removal of N2 and CO2 from CH4 for natural gas upgrading.  相似文献   

19.
Attempts were carried out to enhance the surface hydrophilicity of poly(L ‐lactide), that is, poly(L ‐lactic acid) (PLLA) film, utilizing enzymatic, alkaline, and autocatalytic hydrolyses in a proteinase K/Tris–HCL buffered solution system (37°C), in a 0.01N NaOH solution (37°C), and in a phosphate‐buffered solution (100°C), respectively. Moreover, its chain‐scission mechanisms in these different media were studied. The advancing contact‐angle (θa) value of the amorphous‐made PLLA film decreased monotonically with the hydrolysis time from 100° to 75° and 80° without a significant molecular weight decrease, when enzymatic and alkaline hydrolyses were continued for 60 min and 8 h, respectively. In contrast, a negligible change in the θa value was observed for the PLLA films even after the autocatalytic hydrolysis was continured for 16 h, when their bulk Mn decreased from 1.2 × 105 to 2.2 × 104 g mol?1 or the number of hydrophilic terminal groups per unit weight increased from 1.7 × 10?5 to 9.1 × 10?5 mol g?1. These findings, together with the result of gravimetry, revealed that the enzymatic and alkaline hydrolyses are powerful enough to enhance the practical surface hydrophilicity of the PLLA films because of their surface‐erosion mechanisms and that its practical surface hydrophilicity is controllable by varying the hydrolysis time. Moreover, autocatalytic hydrolysis is inappropriate to enhance the surface hydrophilicity, because of its bulk‐erosion mechanism. Alkaline hydrolysis is the best to enhance the hydrophilicity of the PLLA films without hydrolysis of the film cores, while the enzymatic hydrolysis is appropriate and inappropriate to enhance the surface hydrophilicity of bulky and thin PLLA materials, respectively, because a significant weight loss occurs before saturation of θa value. The changes in the weight loss and θa values during hydrolysis showed that exo chain scission as well as endo chain scission occurs in the presence of proteinase K, while in the alkaline and phosphate‐buffered solutions, hydrolysis proceeds via endo chain scission. © 2002 Wiley Periodicals, Inc. J Appl Polym Sci 87: 1628–1633, 2003  相似文献   

20.
Expressions for the kinetics of grafting chains from a pure end-functional polymer melt to a reactive interface are derived under the assumption that the free energy of reaction is very large and negative. Two cases are considered: a case where the grafting is controlled by diffusion of free end-functional chains through the “brush” of previously grafted chains and a case where the grafting is controlled by the kinetics of the interface reaction itself. Both cases lead to the same form of the grafting kinetics, but with different characteristic times τD = Rga/D, where Rg is radius of gyration, a is the statistical segment length, and D is the diffusion coefficient of the end-functional polymer; and τR = Rg/(akf[B]) where kf is a forward rate constant and [B] is the concentration of groups at the interface that react with the end-function of the polymer chains. The grafting density Σ approaches a practical limit after long times for each value of the degree of polymerization N, and this limit decreases strongly with N. These predictions are compared with the results of recent experiments in which deuterated polystyrene chains with carboxylic acid end groups were grafted to either epoxy- or amine-rich cross-linked epoxy networks. The reaction-controlled grafting model fits the data best and permits us to extract forward rate constants of ˜0.18 kg s−1 mol−1 for the epoxy-rich networks and ˜5 × 10−3 kg s−1 mol−1 for the amine-rich ones.  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