首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 93 毫秒
1.
Three polysilane polymers, (n-PrSiMe) n , (i-PrSiMe) n , and (sec-BuSiMe) n , were synthesized and characterized by DSC. UV spectroscopy, wide-angle X-ray diffraction, and optical microscopy, all at variable temperatures. The known thermochromic transition of (n-PrSiMe) n at 48 C is associated with a change from an orthorhombic to an isotropic phase. (i-PrSiMe) n was examned as an insoluble and soluble (lowM w) fraction, both existing mainly in an orthohombic lattice at room temperature. (sec-BuSiMe) n has a mesophase structure at 25 C, undergoes a weak endothermic transition to a second (nematic) mesophase near 65 C, and becomes isotropic at 160 C.Dedicated to the memory of Professor Zygmunt Lasocki, a fine chemist and a kind and gentle person.  相似文献   

2.
Poly(ethyl-n-butylsilylene) (PEBS) was synthesized by sodium coupling of ethyl-n-butyldichorosilane and separated into fractions with differing molecular weights,M w=1.4×106 and 2.0×104. Both fractions were studied by differential scanning calorimetry and by X-ray diffraction, UV spectroscopy, polarizing optical microscopy, and capillary rheometry, all as a function of temperature. Both samples adopt a hexagonal columnar liquid crystalline structure at room temperature and below. They undergo a weak endothermic transition at ?20°C and a first-order phase transition to a nematic liquid crystalline form at 90°C for the low and 170°C for the highM w fraction. Melting to an isotropic liquid takes place at 106°C for the low and 185°C for the highM w polymer. Both samples undergo two successive thermochromic transitions in the UV, one near the first-order exothermic transition and one near the ?20°C transition; the reasons underlying these thermochromic transitions are discussed. Flow properties of PEBS were investigated as a function of molecular weight.  相似文献   

3.
A series of new alkene and octenyloxy monomers containing 4′-[oligo(ethylene oxide)] n monomethyl ether 4-biphenyl ether carboxyl benzoate [MS3BDBEn] (n = to 3) and 1- (p-methoxydiphenyl)–(carboxyl benzoate) [oligo(ethylene oxide)]n [MSm+2BEnDB] (m = 1, 6; n = 1 to 3) as end groups were synthesized. The molecular structure of the monomers was charaterized using nuclear magnetic resonance (NMR) spectroscopy. These monomers were grafted onto poly(methylhydrosiloxane)s (PMHS) by the platinun-catalyzed hydrosilylation process. The thermal transition temeratures and mesophase textures of the monomers and the polysiloxane polymers have been determined by diffential scanning calorimetry (DSC) and by polarized optical microscopy. The effect of changes in chemical structure on the mesophase properties, glass transition temperature, isotropic temperature, and mesophase texture of the monomers and the polysiloxane polymers is discussed. Polymers PS3BDBEn showed smectic and nematic phases which were not analogous to their precursor nematic monomers MS3BDBEn. Both monomers MSm+2BEnDB and their polymeric homologous PSm+2BEnDB did not exhibit mesophase properties. This demonstrated that the polymer effect could not stabilize the mesophase obtained from mesogenic core which contained a flexible oligo(ethylene oxide) unit interconnecting aromatic group. © 1995 John Wiley & Sons, Inc.  相似文献   

4.
Summary Phase behavior of poly(pyromellitimide)s having (n-alkyloxy)methyl side chains (-CH2OCmH2m+1, m=4, 6, 8) has been studied by wide angle X-ray scattering, differential scanning calorimetry and 13C solid-state NMR. While there are no observable transitions in the polymers bearing short side chain(m=4, 6), there is one transition in the polymer bearing the longest side chain(m=8), which is assigned to mesophase-mesophase transition. All the samples show layered mesophase at room temperature, in which the side chains are amorphous but the main chains form two-dimensional crystals in each layer. In the polymer bearing the longest side chain(m=8), it shows another layered mesophase at high temperature. In the high temperature mesophase, the main chains do not form two-dimensional crystals in each layer; only the lateral packing of the main chains remains undisrupted. Received: 7 July 2000/Accepted: 22 September 2000  相似文献   

5.
The presence of lysine (Lys) or arginine (Arg) and tryptophan (Trp) are important for the antimicrobial effects of cationic peptides. Therefore, we designed and synthesized a series of antimicrobial peptides with various numbers of Lys (or Arg) and Trp repeats [(KW and RW)n-NH2, where n equals 2, 3, 4, or 5]. Antifungal activities of these peptides increased with chain length. Light microscopy demonstrated that longer peptides (n = 4, 5) strongly inhibited in vitro growth of Fusarium solani, and Fusarium oxysporum, at 4–32 μM. Furthermore, longer peptides displayed potent fungicidal activities against a variety of agronomical important filamentous fungi, including F. solani and F. oxysporum, at their minimal inhibitory concentrations (MICs). However, RW series peptides showed slightly higher fungicidal activities than KW peptides against the two strains. Taken together, the results of this study indicate that these short peptides would be good candidates for use as synthetic or transgenic antifungal agents.  相似文献   

6.
Diffusion of pure components (hydrogen (H2), argon (Ar), krypton (Kr), methane (C1), ethane (C2), propane (C3), n-butane (nC4), and n-hexane (nC6)) in silica nanopores with diameters of 1, 1.5, 2, 3, 4, 5.8, 7.6, and 10 nm were investigated using molecular dynamics (MD). The Maxwell-Stefan (M-S) diffusivity (?i,s) and self-diffusivities (Di,self,s) were determined for pore loadings ranging to 10 molecules nm−3. The MD simulations show that zero-loading diffusivity ?i,s(0) is consistently lower, by up to a factor of 10, than the values anticipated by the classical Knudsen formula; the differences increase with increasing adsorption strength. Only when the adsorption is negligible does the ?i(0) approach the Knudsen diffusivity value.MD simulations of diffusion in binary mixtures C1-H2, C1-Ar, C1-C2, C1-C3, C1-nC4, C1-nC6, C2-nC4, C2-nC6, and nC4-nC6 in the different pores were also performed to determine the three parameters ?1,s, ?2,s, and ?12, arising in the M-S formulation for binary mixture diffusion. The ?i,s in the mixture were found to be practically the same as the values obtained for unary diffusion, when compared at the same total pore loading. Also, the ?i,s of any component was practically the same, irrespective of the partner molecules in the mixture. Furthermore the intermolecular species interaction parameter ?12, could be identified with the binary M-S diffusivity in a fluid mixture at the same concentration as within the silica nanopore. The obtained results underline the overwhelming advantages of the M-S theory for mixture diffusion in nanopores.Our study underlines the limitations of the commonly used dusty-gas approach to pore diffusion in which Knudsen and surface diffusion mechanisms are considered to be additive.  相似文献   

7.
Pb0.97La0.03(Zr0.9Ti0.1)0.9925O3 (PLZT 3/90/10) ceramics prepared by solid-state reaction with the compositions near the antiferroelectric/ferroelectric (FE/AFE) phase boundary were studied. From the polarization–electric field P(E) dependence and ex situ X-ray study, an irreversible electric field induced AFE-to-FE phase transition is verified at room temperature. Dielectric and in situ temperature dependent X-ray analysis evidence that the phase transition sequence in PLZT 3/90/10-based ceramics can be readily altered by poling. A first order antiferroelectric-paraelectric (AFE-to-PE) transition occurred at?~190 °C in virgin sample and at?~180 °C in poled sample. In addition, a FE-to-AFE transition occurs in the poled ceramic at much lower temperatures (~120 °C) with respect to the Curie range (~190 °C). The temperature-induced FE-to-AFE transition is diffuse and takes place in a broad temperature range of 72–135 °C. The recovery of AFE is accompanied by an enhancement in the piezoelectric properties.  相似文献   

8.
1H T1 relaxation times were measured in solutions of poly (γ-benzyl-l-glutamate), PBLG, and poly(β-benzyl-l-aspartate), PBLA, and 13C T1 relaxation times, NOE factors and line widths in solutions of the two polypeptides and of poly(l-alanine), PLA, over a broad range of solvent composition (CDCI3—trifluoroacetic acid) including the helix—coil transition. PBLG was also studied through temperature-induced inverted helix—coil transition. Most relaxation parameters change smoothly over the studied range, and the observed changes correspond to increase of mobility with increasing content of trifluoroacetic acid and with increasing temperature except for PBLG which undergoes an inverted helix—coil transition with temperature. Analysis of experimental data by means of the isotropic model yielded, for the backbone carbons of all three polypeptides, τeff ~ 2 × 10?8 sec in the helix and τeff ~ 1 × 10?9 sec in the coil form. For the side-chain carbons, τeff decreases in the range 10?9 ? 10?10 sec with increasing distance from the backbone, with τeff values consistently lower in the coil form. Results of analysis by means of a model with two correlation times are discussed.  相似文献   

9.
The parameters of the electric-field-gradient tensor for copper sites in the HgBa2Ca n ? 1Cu n O2n + 2, Tl2Ba2Ca n ? 1Cu n O2n + 4, and Bi2Sr2Ca n ? 1Cu n O2n + 4 (n = 1–3) lattices have been determined using 67Cu(67Zn) Mössbauer emission spectroscopy and calculated in the framework of the point-charge approximation. The agreement between the experimental and calculated parameters has been achieved under the assumption that the holes formed as a result of the decrease in the oxidation state of a part of the mercury, thallium, or bismuth atoms are distributed over the oxygen sites in the Cu-O or adjacent planes. It has been demonstrated that the oxidation state of cations can be controlled in high-temperature superconducting ceramic materials.  相似文献   

10.
Lihui Cao  Weimin Dong  Xuequan Zhang 《Polymer》2007,48(9):2475-2480
The oxovanadium phosphonates (VO(P204)2 and VO(P507)2) activated by various alkylaluminums (AlR3, R = Et, i-Bu, n-Oct; HAlR2, R = Et, i-Bu) were examined in butadiene (Bd) polymerization. Both VO(P204)2 and VO(P507)2 showed higher activity than those of classical vanadium-based catalysts (e.g. VOCl3, V(acac)3). Among the examined catalysts, the VO(P204)2/Al(Oct)3 system (I) revealed the highest catalytic activity, giving the poly(Bd) bearing Mn of 3.76 × 104 g/mol, and Mw/Mn ratio of 2.9, when the [Al]/[V] molar ratio was 4.0 at 40 °C. The polymerization rate for I is of the first order with respect to the concentration of monomer. High thermal stability of I was found, since a fairly good catalytic activity was achieved even at 70 °C (polymer yield > 33%); the Mn value and Mw/Mn ratio were independent of polymerization temperature in the range of 40-70 °C. By IR and DSC, the poly(Bd)s obtained had high 1,2-unit content (>65%) with atactic configuration. The 1,2-unit content of the polymers obtained by I was nearly unchanged, regardless of variation of reaction conditions, i.e. [Al]/[V], ageing time, and reaction temperature, indicating the high stability of stereospecificity of the active sites.  相似文献   

11.
Stereoisomers of N-(sec-butyl)-N-methyl-N-((R,S)-thiirane 2-ylmethyl) amine ((S)-sec-butyl, optical purity O.P. ~94% and (R)-sec-butyl, O.P. ~26%) were polymerized using a chiral initiator system (1:1) ZnEt2R(-)3,3-dimethyl-1,2-butanediol (R(-)BMBD) (stereoelective polymerization). In agreement with the homosteric character of this initiator, the configuration of the asymmetric centre located in the heterocycle of the elected diastereoisomer is shown to be R from c.d. spectra of residual monomers, whatever the configuration of the asymmetric centre located in the pendant group. O.r.d. and c.d. spectra of stereoelected polydiastereoisomers are described and compared to those of corresponding polydiastereoisomers bearing racemic main-chain chiral centres. 13C n.m.r. spectra of different polydiastereoisomers are reported. The stereosensitivity to tacticity of the main-chain carbons is poor but some of the side-chain carbon atoms are very sensitive to the presence of two asymmetric centres per repeat unit. The optical purity of main-chain chiral centres is deduced from the split resonance of one of these carbons atoms. It is shown that chiroptical properties of polydiastereoisomers primarily depend on configurations of the two chiral sites without obvious contributions from macromolecular conformations. The 242 nm c.d. band assigned to the n → σ1 electronic transition of sulphur chromophores is sensitive only to the optical purity of main-chain chiral centres and can be used to evaluate the configurational enrichment due to the stereoelective polymerization.  相似文献   

12.
Hui Tang  Jiwen Wu  Xinhua Wan 《Polymer》2010,51(23):5482-5489
Raman spectroscopy was applied to study the conformational changes of poly[di(alkyl) vinylterephthalate] (alkyl = n-butyl and sec-butyl). Spectral regions assigned to C-H, C-C and ester carbonyl stretching modes were utilized to provide information on the molecular motion, the relative content of trans and gauche conformers, and the specific interactions existed in the mesophase formation process of these novel thermotropic liquid crystalline polymers. Both the experimental results and the conformational analysis suggested that the aliphatic side groups do have a significant impact on the formation of stable liquid crystalline phase. Based on the van’t Hoff relation, the thermodynamic parameters in the phase transition of poly[di(n-butyl) vinylterephthalate] were estimated (ΔH = 5.52 kJ mol−1, ΔS = 14.02 J mol−1 K−1), which were much smaller than the typical values generally obtained in first-order phase transition. Based on generalized two-dimensional correlation Raman analysis, specific interaction among the phenyl-ring in the mesophase development process was elucidated.  相似文献   

13.
Poly(ethyl-n-butylsilylene) (PEBS) was synthesized by sodium coupling of ethyl-n-butyldichorosilane and separated into fractions with differing molecular weights,M w=1.4×106 and 2.0×104. Both fractions were studied by differential scanning calorimetry and by X-ray diffraction, UV spectroscopy, polarizing optical microscopy, and capillary rheometry, all as a function of temperature. Both samples adopt a hexagonal columnar liquid crystalline structure at room temperature and below. They undergo a weak endothermic transition at –20°C and a first-order phase transition to a nematic liquid crystalline form at 90°C for the low and 170°C for the highM w fraction. Melting to an isotropic liquid takes place at 106°C for the low and 185°C for the highM w polymer. Both samples undergo two successive thermochromic transitions in the UV, one near the first-order exothermic transition and one near the –20°C transition; the reasons underlying these thermochromic transitions are discussed. Flow properties of PEBS were investigated as a function of molecular weight.Presented at the xxvith Silicon Symposium, Indiana University—Purdue University at Indianapolis, March 26–27, 1993.  相似文献   

14.
In this study, silica-based mesoporous materials (the M41S family mesoporous molecular sieves) are synthesized using alkyltrimethylammonium bromide with different chain lengths (CnH2n+1N(CH3)3Br, n = 10, 12, 14, 16) as templates. The resulting silica structures are characterized by X-ray diffraction and are found to exhibit the phase transformation from the hexagonal mesophase MCM-41 to the cubic mesophase MCM-48 (with the space group of Ia3d). The structural phase transition in our study is controlled by the alkyl chain length of the surfactant: with an increase in the surfactant chain length (from C10 to C16), the structure goes from MCM-41 (synthesized by C10), through an intermediate structure (synthesized by C12), to MCM-48 (synthesized by C14 and C16). The amount of ethanol, which is used as a cosolvent, affects the pore size of the structured mesoporous silica, but only to a small extent. In the mean time, the autoclaving time has some effect, though not distinctively, on the structure integrity as well. With increased surfactant to silica ratio, the phase transformation can be shifted to longer chain template.  相似文献   

15.
Nanostructure and morphology and their development of poly(di-n-hexylsilane) (PDHS) and poly(di-n-butylsilane) (PDBS) during the crystal-mesophase transition are investigated using small angle X-ray scattering (SAXS), wide angle X-ray diffraction and hot-stage atomic force microscopy. At room temperature, PDHS consists of stacks of lamellae separated by mesophase layers, which can be well accounted using an ideal two-phase model. During the crystal-mesophase transition, obvious morphological changes are observed due to the marked changes in main chain conformation and intermolecular distances between crystalline phase and mesophase. In contrast to PDHS, the lamellae in PDBS barely show anisotropy in dimensions at room temperature. The nonperiodic structure and rather small electronic density fluctuation in PDBS lead to the much weak SAXS. The nonperiodic structure is preserved during the crystal-mesophase transition because of the similarity of main chain conformation and intermolecular distances between crystalline phase and mesophase.  相似文献   

16.
Poly(p-acryloyloxybenzoic acid) has been obtained in twelve single and two-phase physical states, which include the amorphous glassy and liquid states, the mesomorphic glassy and the mesomorphic liquid states and in addition eight two-phase semicrystalline states (crystal forms I and II). Using mainly differential scanning calorimetry, the transition temperatures, energies and heat capacity changes at the glass transitions have been studied. The time dependency of the glass transition has also been determined. The strongly heating rate dependent amorphous glass transition occurs at 348K (20k min?1 heating rate, ΔCp = 39 JK?1 mol?1), the mesophase glass transition, at 408K (also at 20k min?1, ΔCp = 43 JK?1 mol?1). The latter is less heating rate dependent. The amorphous to mesophase transition occurs between 375 and 475K (ΔH = ?4.5 KJ mol?1); the peaks of melting transitions, which are also strongly heating rate dependent, were observed at 573K and 553K (20 K min?1 heating rate) for crystal forms I and II, respectively. The heat of fusion of crystal form I is estimated to be 22 KJ mol?1. There seems to be no partially amorphous-partially mesomorphic state.  相似文献   

17.
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   

18.
A new nematic-like mesophase photoconductive polymer PPT-TPA consisting of wholly aromatic rigid backbone of poly(p-phenyleneterephthalate), PPT and pendent hole-transporting triphenylamine (TPA) groups attached to the ends of oxydecyl spacers has been synthesized. The photorefractive composite contains the photoconductor PPT-TPA, the chromophore diethylaminodicyanostyrene (DDCST), and the photosensitizer C60. Although no plasticizer was added, the glass transition temperature Tg of the composite is 15 °C, which characterizes it as a low-Tg photorefractive material. We investigate the correlation between the mesophase structure and its optical/physical properties by X-ray diffraction, photoconductive and photorefractive experiments. The new composite and its properties are compared to PPT-CZ composites with only a different charge transporting agent (carbazole, CZ) but a much more ordered mesophase structure, which were studied previously and have shown very good photorefractive properties. Despite of a lower photoconductivity of the new photorefractive composite PPT-TPA (n=10):DDCST:C60 this material shows a higher photorefractive sensitivity Sn2 of 2±0.2 cm2/kJ at E=50 V/μm than the previously synthesized composite PPT-CZ (n=10):DDCST:C60.  相似文献   

19.
(K, Na)NbO3-based lead free materials have been found to exhibit good piezoelectric properties due to the orthorhombic–tetragonal polymorphic phase transition (PPT) temperature compositionally shifted downward to near room temperature. However, this transition correspondingly results in a strong temperature dependence of the dielectric and piezoelectric properties. In this work, new quaternary (1?x) (K0.4425Na0.52Li0.0375)(Nb0.8925Sb0.07Ta0.0375)O3 (KNLNST)–xSrTiO3 (ST) lead-free piezoelectric ceramics were fabricated by a conventional ceramic technique and their structure and piezoelectric properties were also studied. The results of X-ray diffraction reveal that SrTiO3 diffuses into the KNLNST lattices to form a new solid solution with a perovskite structure. After the addition of SrTiO3, tetragonal–orthorhombic phase transition shifts to lower temperatures. The good piezoelectric properties of 0.995 KNLNST–0.005 ST material were found to be d33~295 pC/N, kp~42%, and εr~1902, with greatly improved temperature stability over the temperature range of 0–100 °C, demonstrating practical potential for actuator and ultrasonic transducer applications.  相似文献   

20.
《Ceramics International》2022,48(2):1879-1888
In the broad application of tunable devices, the temperature stability of the dielectric tunable performance of ceramics should be considered. In this work, (1-x)Ba(Zr0.36Ti0.64)O3x(Ba0.82Ca0.18)TiO3 (BZ0.36T–xBC0.18T) ceramics, where x = 0.3–0.7, were prepared by the solid-state high-temperature method. The combined effects of Ca2+ and Zr4+ ions on the microstructure of BZ0.36T–xBC0.18T ceramics were observed. The crystal structure was analyzed through X-ray diffraction, showing a pure perovskite ABO3-type structure, indicating the formation of a BCZT solid solution. High dielectric tunability (nr > 85% at < 10 kV/cm) under a low DC bias field is achieved in BZ0.36T–0.5BC0.18T and BZ0.36T–0.6BC0.18T, especially the maximum value of nr ~87.51% at 7.68 kV/cm obtained for BZ0.36T–0.6BC0.18T. High tunability under a low DC bias field may be related to not only the extrinsic contribution (dipole reorientation, domain-wall motion) but also the intrinsic contribution (lattice phonon or anharmonic interactions of B-site ions). The maximum FOM value of ~847 is achieved in BZ0.36T–0.5BC0.18T, which is related to the high nr ~74.57% and lower tan δ ~8.8 × 10?4 at 400 V, demonstrating its excellent performance for tunable device applications. Furthermore, the FOMT curve of BZ0.36T–0.3BC0.18T ceramics in the range of 126–269 is flatter than other compositions in the temperature range of ?20°C-100 °C, showing an improved dielectric tunable performance with better temperature stability. The improved temperature stability of BZ0.36T–0.3BC0.18T may contribute to the enhancement of the diffuse phase transition (DPT) degree, which results in a flattened phase transition region. These results suggest that BZ0.36T–0.3BC0.18T is competitive candidate for temperature-stable tunable device applications, and the compositional tailored DPT can be expected to be a feasible means to improve the temperature stability of dielectric tunable performance.  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