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 296 毫秒
1.
We have measured the time evolution of the self‐assembly process in perpendicular‐oriented cylindrical‐phase diblock copolymer thin films using statistical analysis of high‐resolution scanning electron microscope (SEM) images. Within minutes of annealing above the polymer glass‐transition temperature, microphase separation between polymer blocks results in formation of uniform nanometer‐scale domains whose relative position is initially largely uncorrelated. On further annealing, the cylindrical polymer domains organize into a two‐dimensional hexagonal lattice whose characteristic grain size increases slowly with time (~t1/4). © 2004 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 42: 1970–1975, 2004  相似文献   

2.
The microstructure of polymer matrix under cylindrical confinement is key to understanding the size‐dependent thermomechanical behavior of electrospun nanofibers. Coarse‐grained molecular dynamics simulation was applied here to probe polymer systems under cylindrical confinement, prepared with or without pre‐stretching. Simulation results showed that below a certain radius, a noticeable increase of the elastic modulus is observed with the decrease of the radius of cylindrical confinement. This size‐dependent mechanical behavior correlated to the degree of polymer chain orientation. Modulation of density and bond orientation in the radial direction was observed: the density and bond orientation began to oscillate, increasing the oscillation amplitudes with decreases in the radius. Such behavior suggests that the cylindrical confinement enhances the bond alignment of the entire fiber and not in the near‐surface layers only. The unstretched fibers had uniform density distribution along the fiber axis, while the stretched fibers demonstrated a fluctuation in density distribution. The crossover radius of size‐dependent behavior was two orders of magnitude smaller than observed in real experiments, demonstrating that the confinement affects some internal fiber scale, which exceeds the scale of individual macromolecules, and this internal scale may be related to supramolecular structures of the polymer matrix rather than the individual macromolecules. © 2017 Wiley Periodicals, Inc. J. Polym. Sci., Part B: Polym. Phys. 2017 , 55, 506–514  相似文献   

3.
In this study, we present nanodomes‐combined surface relief gratings (SRGs) of azopolymer films with controlled shapes and sizes. We investigate the effect of the polarization mode of light interference on leading nanodomes in the conventional SRG patterns. In addition, we also systematically study the relationship between Bragg distance of light interference and shapes of nanodomes. From this, we explain the anisotropic self‐assembled behavior nanodomes in photoaddressable azopolymer films regarding polarization modes as well as spatial confinement effect. © 2019 Wiley Periodicals, Inc. J. Polym. Sci., Part B: Polym. Phys. 2019 , 57, 731–737  相似文献   

4.
Although there have been many reports on the preparation and applications of various polymer nanofibers with the electrospinning technique, the understanding of synthetic parameters in electrospinning remains limited. In this article, we investigate experimentally the influence of solvents on the morphology of the poly(vinyl pyrrolidone) (PVP) micro/nanofibers prepared by electrospinning PVP solution in different solvents, including ethanol, dichloromethane (MC) and N,N‐dimethylformamide (DMF). Using 4 wt % PVP solutions, the PVP fibers prepared from MC and DMF solvents had a shape like a bead‐on‐a‐string. In contrast, smooth PVP nanofibers were obtained with ethanol as a solvent although the size distribution of the fibers was somewhat broadened. In an effort to prepare PVP nanofibers with small diameters and narrow size distributions, we developed a strategy of using mixed solvents. The experimental results showed that when the ratio of DMF to ethanol was 50:50 (w/w), regular cylindrical PVP nanofibers with a diameter of 20 nm were successfully prepared. The formation of these thinnest nanofibers could be attributed to the combined effects of ethanol and DMF solvents that optimize the solution viscosity and charge density of the polymer jet. In addition, an interesting helical‐shaped fiber was obtained from 20 wt % PVP solution in a 50:50 (w/w) mixed ethanol/DMF solvent. © 2004 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 42: 3721–3726, 2004  相似文献   

5.
Ultrahigh‐molecular‐weight linear polyethers were prepared through a reaction between the phenylquinoxaline monomers 2,3‐bis(4‐hydroxyphenyl)‐6‐fluoroquinoxaline and 2,3‐bis(4‐hydroxyphenyl)‐6‐(α,α,α‐trifluoromethyl)quinoxaline and 1,12‐dibromododecane. A new hyperbranched polyether containing a phenylquinoxaline moiety was also prepared from a new self‐polymerizable AB2 monomer, 2,3‐bis(6‐bromohexyloxyphenyl)‐6‐(4‐hydroxyphenyloxy)quinoxaline. All the polyethers were amorphous and soluble in polar aprotic solvents. Their solution‐cast thin films were light yellow, ductile, and optically transparent. The polymers were thermally stable up to 350 °C and had glass‐transition temperatures in the range of 25–83 °C, which depended on the architecture and monomer structure. The monomers and polymers displayed fluorescence maxima in the blue‐light region in the range of 431–449 nm with relatively narrow peak widths; this indicated that they had pure and intense fluorescence. © 2004 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 42: 3587–3603, 2004  相似文献   

6.
A hydrodynamic scattering treatment of interacting polymer chains is extended to obtain the five‐point chain–chain–chain–chain–chain hydrodynamic interaction tensor. The tensor is used to calculate the second‐order concentration correction to the self‐diffusion coefficient of a polymer in solution. The self‐similarity assumption of the hydrodynamic scaling model of polymer dynamics is tested against these calculations. © 2004 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 42: 1663–1670, 2004  相似文献   

7.
We report the self‐consistent field theory (SCFT) of the morphology of lamella‐forming diblock copolymer thin films confined in two horizontal symmetrical/asymmetrical surfaces. The morphological dependences of thin films on the polymer‐surface interactions and confinement, such as film thickness and confinement spatial structure, have been systematically investigated. Mechanisms of the morphological transitions can be understood mainly through the polymer‐surface interactions and confinement entropy, in which the plat confinement surface provides a surface‐induced effect. The confinement is expressed in the form of the ratio D/L0, here D is film thickness, and L0 is the period of bulk lamellar‐structure. Much richer morphologies and multiple surface‐induced morphological transitions for the lamella‐forming diblock copolymer thin films are observed, which have not been reported before. © 2008 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 47: 1–10, 2009  相似文献   

8.
Random THV terpolymers consisting of tetrafluoroethylene (TFE), hexafluoropropylene (HFP), and vinylidene fluoride (VDF) are viable alternatives to polytetrafluoroethylene (PTFE) combining excellent chemical stability and thermoplastic processability. Although the properties of THV may be modified by crystallization, little is known on how crystallization is influenced by the chain microstructure of THV. We analyzed the chain microstructure of THV‐221G by solid‐state 19F NMR spectroscopy under fast magic angle spinning, revealing that THV‐221G contains 43.8 mol % TFE, 46.0 mol % VDF, and 10.2 mol % HFP. Sequence analysis revealed that the TFE units are preferentially located next to other TFE units. The HFP units, which are obstacles to crystallization because of their bulky CF3 side groups, are preferentially located next to VDF units. WAXS measurements correspondingly revealed the presence of THV‐221G crystals with PTFE‐like packing and of further THV‐221G crystal populations with widened d‐spacings caused by the incorporation of certain amounts of HFP units into the THV‐221G crystals. Under confinement imposed by the cylindrical nanopores of self‐ordered alumina, the THV‐221G melting point decreased with decreasing pore diameter. Although direct impingement of the growing THV‐221G crystals on the pore walls is unlikely, the geometric confinement limits the access of growing THV‐221G crystals to crystallizable THV‐221G chain segments. © 2019 Wiley Periodicals, Inc. J. Polym. Sci., Part B: Polym. Phys. 2019 , 57, 1402–1408  相似文献   

9.
AB diblock copolymers dissolved in a selective solvent for block B spontaneously self‐assemble into spherical or cylindrical micelles with an A core surrounded by a corona of the B block. Using a Monte Carlo lattice model, we find that the sphere‐to‐cylinder transition can be fully described in terms of a packing parameter that depends on the relative cross sections of the two blocks. The latter can be varied either by a block length being changed or by the B block being allowed to adopt an ordered conformation (e.g., helical). The model results are in good agreement with experimental observation. They also shed some light onto the molecular origins of the ease of processing and exceptional mechanical properties of spider dragline. © 2002 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 40: 890–895, 2002  相似文献   

10.
In this article we report a study of in situ polymerization of ethylene by intercalated montmorillonite (MMT) with metallocene, allowing an investigation of the nanoscopic confinement effect of olefin polymerization and of the structure of polymer prepared in situ. Ethylene polymerization by intercalated MMT with metallocene and the varied aggregation morphology of the resulting polymer during polymerization were studied by X‐ray diffraction (XRD), differential scanning calorimetry (DSC), and gel permeation chromatography (GPC). The polymerization kinetics and the resulting polymer before and after destruction of the silicate registry were different. The laminated structure of silicate lowered the all‐reaction rate, including the propagation, chain transfer, and termination reactions, producing polymer of a high molecular weight. Moreover, the melting point of the polymer gradually increased during the in situ polymerization, indicating that nanoscopic confinement between solid surfaces affects the crystallization behavior of polyethylene via in situ polymerization. © 2003 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 42: 38–43, 2004  相似文献   

11.
The self‐assembly of a metallo‐supramolecular PS‐[Ru]‐PEO block copolymer, where ‐[Ru]‐ is a bis‐2,2′:6′,2″‐terpyridine‐ruthenium(II) complex, in thin films was investigated. Metallo‐supramolecular copolymers exhibit a different behavior as compared to their covalent counterparts. The presence of the charged complex at the junction of the two blocks has a strong impact on the self‐assembly, effecting the orientation of the cylinders and ordering process. Poly(ethylene oxide) cylinders oriented normal to the film surface are obtained directly regardless of the experimental conditions over a wide range of thicknesses. Exposure to polar solvent vapors can be used to improve the lateral ordering of the cylindrical microdomains. © 2008 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 46: 4719–4724, 2008  相似文献   

12.
We demonstrate the ability of the reversible addition‐fragmentation chain transfer (RAFT) process to produce well‐defined block co‐oligomers for which each block has a narrow molecular weight distribution and degrees of polymerization ranging from 2 to 33. We exploit RAFT versatility to control the structure of the co‐oligomers and produce amphiphilic block co‐oligomers of styrene, acrylic acid and ethylene glycol. A detailed study shows that the amphiphilic diblock co‐oligomers self‐assemble in solution and form micelles or particles, depending on the hydrophobicity of the diblock. These oligomers present an excellent alternative to traditional amphiphilic molecules, by combining the properties of polymers with those of single molecule surfactants. © 2011 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2012  相似文献   

13.
A reversible catalyst immobilization system via self‐assembly of hydrogen bonding between thymine anchored on silica gel support and 2,6‐diaminopyridine functionalized with a catalyst (copper bromide‐N,N,N′,N′‐tetraethyldiethylenetriamine (TEDETA) complex) was developed for the atom transfer radical polymerization (ATRP) of methyl methacrylate (MMA). At elevated temperatures, the hydrogen bonding disassociated and released the catalyst as free small molecules for catalysis, which effectively mediated a living polymerization of MMA, producing PMMA with controlled molecular weight and narrow molecular weight distribution (<1.3). At room temperature, the catalyst assembled on the silica gel support by hydrogen bonding, and thus could be recovered and reused for a second run of ATRP. The recovered catalyst still mediated a living polymerization of MMA with reduced activity (54–64%), but had much improved control of the polymerization. The resulting PMMA had molecular weights very close to theoretical vales. © 2003 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 42: 22–30, 2004  相似文献   

14.
The metal‐catalyzed living radical polymerization of methyl methacrylate, n‐butyl acrylate, and styrene, initiated with p‐toluenesulfonyl bromide and phenoxybenzene‐4,4′‐disulfonyl bromide and catalyzed with CuBr/2,2′‐bipyridine (bpy) and various self‐regulated Cu‐based catalytic systems such as Cu2O/bpy, Cu2S/bpy, Cu2Se/bpy, and Cu2Te/bpy, is reported. Similarities and differences between the arenesulfonyl chloride and arenesulfonyl bromide initiators are discussed. The arenesulfonyl bromide initiators require reduced reaction times to produce polymers in high conversions under milder reaction conditions than the corresponding arenesulfonyl chloride initiators. At the same time, they exhibit 100% initiator efficiency and generate polymers with narrow molecular weight distributions and functional chain ends. © 2004 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 43: 319–330, 2005  相似文献   

15.
We established a self‐catalyst system for solid phase incorporation of gaseous carbon dioxide into terpolymers prepared by polymerization of glycidyl methacrylate, N‐benzyl‐N‐[2‐(methacryrolroxy)ethyl]‐N,N‐dimethylammonium bromide, and methyl methacrylate. Terpolymer composition affected the incorporation behavior where the terpolymer with higher oxirane content exhibited higher efficiency of carbon dioxide incorporation. © 2004 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 42: 4941–4947, 2004  相似文献   

16.
Particulate polymer‐in‐polymer mezodispersions show a pronounced increase in the size of the dispersed particles during melt‐phase annealing. Three ripening mechanisms have been proposed: Brownian coalescence, Ostwald ripening, and hydrodynamic coarsening. The modified Cahn–Hilliard equation predicts growth by Ostwald ripening and diffusion‐induced coalescence. Simulations of this mechanism show a self‐similar particle size distribution, but the distribution broadens with the increasing volume fraction of the minor phase. Hydrodynamic coarsening caused by concentration gradients and random Brownian forces has been simulated according to the hydrodynamic model. The simulations show that concentration‐driven hydrodynamics have little effect on the particle size distribution. Experiments have been performed to investigate the relative importance of these ripening mechanisms for polybutadiene in a polystyrene system. © 2004 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 42: 603–612, 2004  相似文献   

17.
Multilayer self‐assembled films, consisting of partially doped polyaniline (PAN) as a polycation and water‐soluble poly(4‐carboxyphenyl)acetylene (PCPA) as a polyanion, were fabricated through electrostatic attraction. These ultrathin PCPA/PAN films were relatively stable toward aqueous electrolyte solutions and polar organic solvents, and the photoelectric conversion properties of the self‐assembled PCPA/PAN films could be measured with traditional three‐electrode cells in 0.5 M KCl aqueous solutions. With an increase in the bilayer number, the photocurrent rose, and it reached its maximum at eight bilayers. A further increase led to a current drop due to an increase in the recombination probability and weak visible‐light transmission. When neutral PAN films were used, the photocurrent increased consecutively within 15 bilayers, and this indicated that the PAN molecules in the neutral state were stronger electron donors than those in the partially doped state. © 2004 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 42: 3224–3229, 2004  相似文献   

18.
For a range of applications, polymers are now being patterned into nanometer‐sized features. In these applications, the robust mechanical properties of the nanostructures are critical for performance and stability. Brillouin light scattering is presented as a nondestructive, noncontact tool used to quantify the elastic constants in such nanostructures. We demonstrate this through a series of thin films and parallel ridges and spacings (gratings) with ridge widths ranging from 180 to 80 nm. For the set of films and structures presented here, the room‐temperature elastic moduli did not change with decreasing film thickness or grating ridge width, and this implied that one‐dimensional and two‐dimensional confinement‐induced changes of the mechanical properties were not significant down to feature sizes of 80 nm. Additionally, Brillouin spectra of submicrometer gratings revealed new modes not present in the spectra of thin films. The origin of these new modes remains unclear. © 2004 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 42: 1106–1113, 2004  相似文献   

19.
We report on a self‐organized striped pattern of refractive indices in a photopolymer film. Silicon oxide (SiO2) particles (1.3‐μm diameter) were dispersed on a cover glass above a monomer mixture (bifunctional methacrylate monomer, cyanoethylmethacrylate, and 2,4,6‐trimethylbenzoyldiphenylphosphine oxide), and the sample was uniformly irradiated from a tube light source. The self‐organized striped pattern was observed inside a 0.45‐mm thickness photopolymer film. In contrast, no pattern was observed in a 0.45‐mm thickness sample photopolymerized without particles on a cover glass. The results suggest that the SiO2 particles acted as “lenses” and produced a light distribution reflecting the shape of the light source used during the irradiation. The photopolymerization extended in depth according to the light distribution, and then the striped pattern was generated inside a medium. The photopolymer film with the striped pattern has an angular dependence in the transmittance as a result of the microstructure and can be applied to a flexible optical transmission filter. © 2004 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 42: 3351–3358, 2004  相似文献   

20.
The self‐assembly of head‐tail type block copolymers composed of polyamidoamine dendron head block and poly(L ‐lysine) (PLL) tail block was studied using a light scattering technique and transmission electron microscopy. A PLL tail block in a head‐tail type block copolymer exhibits a coil‐to‐helix transition as a result of the change in solvent quality from water to methanol. When the PLL tail block takes a helical conformation in high methanol content, the resulting head‐tail type block copolymer has a defined three‐dimensional structure like that of a protein molecule. Self‐assemblies of such block copolymers having a totally fixed molecular shape spontaneously form polymersome‐like self‐assemblies with an extremely narrow size distribution through converging to a thermodynamically stable assembling state. © 2009 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 47: 1217–1223, 2009  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